Download The Urinary Tract

Document related concepts

Molecular mimicry wikipedia , lookup

Biofilm wikipedia , lookup

Marine microorganism wikipedia , lookup

Horizontal gene transfer wikipedia , lookup

Triclocarban wikipedia , lookup

Hospital-acquired infection wikipedia , lookup

Trimeric autotransporter adhesin wikipedia , lookup

Magnetotactic bacteria wikipedia , lookup

Bacterial cell structure wikipedia , lookup

Human microbiota wikipedia , lookup

Urinary tract infection wikipedia , lookup

Bacterial morphological plasticity wikipedia , lookup

Transcript
Stony Brook University
The official electronic file of this thesis or dissertation is maintained by the University
Libraries on behalf of The Graduate School at Stony Brook University.
©
©A
Allll R
Riigghhttss R
Reesseerrvveedd bbyy A
Auutthhoorr..
Role of Type II Secretion and Type IV pilus Systems in
the Virulence of Uropathogenic Escherichia coli
A Dissertation Presented
by
Ritwij Kulkarni
to
The Graduate School
in Partial Fulfillment of the
Requirements for
the Degree of
Doctor of Philosophy
in
Genetics
Stony Brook University
December 2007
Stony Brook University
The Graduate School
Ritwij Kulkarni
We, the dissertation committee for the above candidate for the
Doctor of Philosophy Degree, hereby recommend
acceptance of this dissertation.
Dr. David G. Thanassi- Dissertation Advisor
Associate Professor
Department of Molecular Genetics and Microbiology
Dr. James B. Bliska- Chairperson of Defense
Professor, Department of Molecular Genetics and Microbiology
Dr. Jorge L. Benach
Director, Center for Infectious Diseases
Professor and Chair, Department of Molecular Genetics and Microbiology
Dr. Matthew A. Mulvey
Associate Professor, Department of Pathology
University of Utah
This Dissertation is accepted by the Graduate School
Lawrence Martin
Dean of Graduate School
ii
ABSTRACT OF DISSERTATION
Role of Type II Secretion and Type IV pilus Systems in
the Virulence of Uropathogenic Escherichia coli.
by
Ritwij Kulkarni
Doctor of Philosophy
in
Genetics
Stony Brook University
2007
Type II secretion systems (T2SS) and the evolutionarily related type IV pili (T4P) are
important virulence determinants in many Gram-negative bacterial pathogens.
However, the roles of T2SS and T4P in the virulence of extraintestinal pathogenic
Escherichia coli have not been determined. Extraintestinal pathogenic E. coli cause a
variety of diseases, including urinary tract infections and neonatal meningitis. We
investigated the function of T2SS and T4P in the model uropathogenic E. coli
(UPEC) strains UTI89 and CFT073, which possess T2SS and T4P gene clusters
homologous to clusters present in laboratory strains of E. coli. In addition, UTI89
contains a second T2SS gene cluster homologous to one present in enterotoxigenic E.
coli (ETEC) strains that cause diarrhea. We constructed deletion mutations to disable
iii
each T2S and T4P system in the UPEC strains, and compared these mutants with
their wild type counterparts using tissue culture assays and the CBA/J mouse model
of ascending urinary tract infection. No deficiencies were observed with any of the
mutants in adherence, invasion or replication in human bladder or kidney cell lines,
but UTI89 ∆hofQ (T4P) and UTI89 ∆gspD (ETEC-T2SS) exhibited approximately 2fold defects in fluxing out of bladder epithelial cells. In the mouse infection model,
each of the knockout mutants was able to establish successful infections in the
bladder and kidneys by day one post-infection. However, UTI89 ΔhofQ and a
CFT073 ∆hofQ ∆yheF (K12-T2SS) double mutant both exhibited defects in
colonizing the kidneys by day seven post-infection. Furthermore, the plasmid-borne
copies of hofQ and/or yheF were able to complement the kidney colonization defect
of the deletion mutants. These results establish T4P and potentially also secreted
T2SS effectors as virulence determinants of UPEC important for persistence in the
urinary tract, particularly in renal tissues.
iv
Dedication
To my grandparents,
Mrs. Tara and Prof. Vasant V. Upadhye.
“I owe so much of this to you…”
Table of Contents
List of Figures and Tables..…………………………………………………….…..vii
Acknowledgements…….....…………………………………………………….….viii
Chapter One: Introduction…………………………………………………………...1
I.
The Urinary Tract…………………………………………………………...1
II. Uropathogenic Escherichia coli………………………………………….….2
III. Virulence Factors Produced by the UPEC……………………………….….3
IV. Secretion Systems in the Gram-Negative Bacteria………………………….5
V. Type II Secretion System……………………………………………………7
VI. Type IV Pilus System……………………………………………………...10
VII. Histology of the Urinary Tract………………………………………….….13
VIII. UPEC Developmental Cascade and Urinary Tract Pathogenesis………….14
IX. Recurrent UTI (rUTI)……………………………………………………...17
X. Host Responses to the Infection by UPEC………………………………...18
XI. Model for UPEC Infections in the Bladder...………………………………20
XII. Model Organisms……………………………………………………….….21
XIII. Figures..…………………………………………………………………….23
Chapter Two: Materials and Methods………………………………………………28
Table 2.1 List of strains from panel of pathogenic and non-pathogenic E. coli….….38
Table 2.2 List of primers used in this study..………………..……………………….40
Chapter Three: T2S and T4P Systems Contribute to the Virulence of UPEC.…….43
I.
Introduction……………………………………………………………..….43
II.
Results………………………………………………………………….. …45
III.
Discussion………………………………………………………………….50
IV.
Figures………………………………………………………………….…..53
Chapter Four: Investigation of the Molecular Basis for the Virulence Defects of
UTI89 ΔhofQ and CFT073 ΔyheF, ΔhofQ..…………………………………………61
Introduction………………………………………………………….….………..61
Results…………………………………………………………………..………..63
Discussion………………………………………………………………………..68
Figures………………………………………………………………….………...70
Chapter Five: Discussion and Future Directions….…………………………….….79
I.
Conclusions………………………………………………………………...79
II. Future Directions...………………………………………………………...84
Abbreviations..………………………………………………………………………..89
References……………………………………………………………………………..90
vi
List of Figures and Tables
Figure 1.1 Human urinary tract…..….………………………………………………23
Figure 1.2 Schematic representation of sec-independent and sec-dependent
pathways..………………………….………………………………………………24
Figure 1.3 Schematic representation of type II secretion and type 4 pilus
biogenesis.………………………….………………………………………………25
Figure 1.4 Progression of bladder infection by UPEC………………………………26
Figure 1.5 UPEC developmental cascade in the host bladder…..……………….….27
Table 2.1 List of strains from panel of pathogenic and non-pathogenic E.coli.....….38
Table 2.2 List of primers used in this study………………..…………………….….40
Figure 2.1 Schematic representation of λRed recombinase-based gene deletion ……42
Figure 3.1 Schematic representation of the T2SS and T4P gene clusters in UTI89 and
CFT073.…………………………….……………………………………………...53
Figure 3.2 Expression of the T2SS and T4P secretin genes…………………...……54
Figure 3.3 PCR verification of the different secretin knockout mutants……………55
Figure 3.4 Comparison of growth curves of the OM secretin deletion mutants
and their WT counterparts ……………….…………………………………..……56
Figure 3.5 Mouse infection experiments.……………………………………………57
Figure 3.6 Complementation of the UTI89 and CFT073 mutants …………...….….58
Figure 3.7 Histological examination of the kidney sections from mice infected with
mutant and WT strains of UTI89…………………………………..………………59
Figure 3.8 Histological examination of the kidney sections from mice infected with
mutant and WT strains of CFT073………………………………..……………….60
Figure 4.1 Adhesion of secretin deletion mutants and WT strains to bladder and
kidney epithelial cell lines.……………………………………………..………….70
Figure 4.2 Invasion of secretin deletion mutants and WT strains into bladder and
kidney epithelial cell lines..…………………………………………..……………71
Figure 4.3 Efflux of WT UTI89 and the indicated UTI89 secretin mutants from 5637
human bladder epithelial cells..…………………………………..………………..72
Figure 4.4 Intracellular replication of secretin deletion mutants and parental WT
strains of UTI89…………….…………………………………………..………….73
Figure 4.5 Instant cytotoxicity of UTI89 (A) or CFT073 (B) for human bladder
epithelial cells……………………....…………………………………..………….74
Figure 4.6 Cytotoxicity caused by UTI89 during the 6 h efflux period.....…………75
Figure 4.7 Efflux index values for the representative ETEC-T2SS negative
strains……………………….…………………………………………..………….76
Figure 4.8 Role of T4P in the adhesion to human bladder and kidney epithelial
cells………………………………....…………………………………..………….77
Figure 4.9 Biofilm Formation………………………………………........………….78
vii
Acknowledgements
In the past few years as a graduate student, I became friends with many people. Their
whole hearted encouragements and timely rebukes have made this research endeavor and
my survival in the US possible as well as fun.
I am deeply indebted to my thesis mentor Dr. David Thanassi. The reason I decided
to join David’s lab was his affable and optimistic nature. His positive attitude always
meant a lot to me and in future that’s something I am going to miss. Furthermore, David
is a mentor in a true sense of the word. All Ph.D. advisors impart some scientific
knowledge on their students. That’s part of their job. But David took this idea of
mentorship beyond the lab. Once he actually taught me how to change a flat tire on my
car. In Utah, he taught me how to ski. In a restaurant, when I got grapefruit with my
breakfast menu, he even taught me how to eat it. David has been my mentor both inside
and outside the lab and I am certainly going to miss him.
I would also like to thank my committee members, Dr. James Bliska, Dr. Jorge
Benach and Dr. Matthew Mulvey (Matt) for their insights, and all the simple and difficult
questions that they asked me through the course of this project.
Matt is also our collaborator for this project. David and I went to Salt Lake City to his
lab where Matt and his graduate student Bijaya K. Dhakal, taught me how to catheterize
mice. They also carried out some of the initial mouse infection experiments. I thank them
for being excellent troubleshooters of anything mice! I also thank Matt and his family for
being the gracious hosts they were during our stay in Utah.
Catheterizing mice on my own in Stony Brook was a nerve wrecking experience. If
Ms. Gloria Monsalve were not there to hold my hand and guide me patiently, I could not
have done a single mouse experiment. I thank Gloria and Mr. Patricio Mena for their
help in mouse work.
I became friends with Ms. Nadine Henderson, Ms. Dreania Levine and Ms. Kate Bell
because of our shared interests in Harry Potter, movies, literature and/or politics. I thank
them for their love and friendship.
Dr. Tony Ng, my best friend and former apartment mate, stood by me through ups
and downs; always helping and advising in all matters scientific and non-scientific. Tony
and other friends such as Ms. Zankhana Thakkar, Dr. Subhra Chakraborty and Mr.
Zachary Kurtz made my graduate years and the Stony Brook experience memorable.
Ms. Jen Parla nee Choy and I are batchmates. We faced similar problems through the
course of graduate school and hence were able to share our sorrows to reduce the pain. I
will always cherish Jen’s advice and how she would make me see the silver linings of the
darkest of the clouds.
The support and friendship and all the gossip that we shared in our morning coffee
group that included Mary, Betty and Kate (Kline), helped me survive the last two years of
graduate school.
My parents, Ms. Mrinal and Mr. Purushottam Kulkarni have been a driving force in
my life. Whatever I am today is because of them
And finally my wife, Ms. Priyadarshini Pathak (Priya): without her dedication, love
and encouragement, I could not have faced the challenges both inside and outside the lab.
- RUDY (Ritwij Kulkarni)
Chapter One
Introduction
I. The Urinary Tract
The human urinary tract is a multi-organ complex consisting of the kidneys,
ureters, urinary bladder and urethra; the organs involved in production, storage and
excretion of urine. Blood is filtered in the glomerular filters of the kidneys at a rate of 125
ml/min (44). The glomerular filtrate consists of glucose, excess salts, urea and mainly
water. Ninety-nine percent of water in the glomerular filtrate is reabsorbed into the blood
to leave behind the urine. Urine is carried by the two ureters into the urinary bladder,
where it is stored for excretion through the urethral opening (Fig. 1.1).
Since the urinary tract is open to the external environment, it is easy for pathogens
to gain entry and establish infection. To prevent this, there are many innate defenses; for
example, the gush of urine purges the urinary tract of any non-adherent particles as well
as pathogens, the low pH and osmolarity of urine has a bacteriostatic effect and the
sphincters at the bladder end of the ureters prevent back-flow of urine into the kidneys.
These defense mechanisms can be breached by different pathogenic microorganisms to
cause urinary tract infections (UTIs), including cystitis (infection of bladder) and
pyelonephritis (infection of kidney). A descending UTI occurs as a consequence of
bacteremia. Blood-borne bacteria enter the kidneys through the glomerular filters, an
event more likely in neonates due to immaturity of the filters. An ascending UTI occurs
due to the entry of the bacteria in the vicinity of the urethral opening into the urinary
1
tract, primarily due to fecal contamination. If left untreated, the pathogens may ascend
the ureters into the kidneys establishing acute pyelonephritis. Presumably, because of
their longer urethrae, males are less prone to ascending UTIs than females.
II. Uropathogenic Escherichia coli
In 1885, a German pediatrician and bacteriologist Dr. Theodore Escherich
discovered the bacteria that were causative agents of infantile diarrhea and gastroenteritis
(90). Later in 1919, these bacteria were grouped under the genus Escherichia. Most
members of this genus colonize the lower intestines of mammals within a few hours after
birth as commensals, with E. coli as the principal commensal species. These Gramnegative, facultative anaerobic, motile rod shaped bacteria belong to the γ suborder of the
family Proteobacteriaceae. They cause disease rarely, except in immunocompromised
individuals or where the normal gastrointestinal barriers are breached. There is a great
diversity of E. coli strains in the environment, some of which are known to infect human
beings while the others such as E coli K12 are commonly used research tools. Diseases
caused by pathogenic E. coli in humans can be divided into three general clinical
syndromes: enteric/diarrheal disease, UTIs, and sepsis/meningitis (52). The pathogenic E.
coli causing UTI (uropathogenic E. coli; UPEC) and sepsis/meningitis (Neonatal
meningitis associated E. coli; NMEC) in humans, and respiratory infections, pericarditis
and peritonitis in birds (avian pathogenic E. coli; APEC), belong mainly to phylogenetic
group B2 (intestinal pathogenic E. coli belong to groups A, B1 or D (49)) and exhibit a
considerable overlap with respect to virulence factors and serotypes (21, 52). Owing to
2
these similarities, the UPEC, NMEC and APEC are grouped under an inclusive
designation: extra-intestinal pathogenic E. coli (ExPEC) (83).
UTIs are of great interest because in North America alone, 40% of women suffer
from a UTI at least once during their lifetime (25), leading to 7 million office visits every
year with total monetary costs approaching $2 billion (27). The causative agents of UTIs
come from different groups of microorganisms such as fungi (Candida albicans), Gramnegative (E. coli, Proteus mirabilis, Pseudomonas aeruginosa) and Gram-positive
(Enterococcus faecalis, Staphylococcus saprophyticus) bacteria. Though these causative
agents show a different distribution depending on whether the infection is hospital- or
community-acquired, UPEC is the major culprit, causing 80% of the hospital- and 40%
of the community-acquired infections (82).
The success of UPEC as pathogens can be attributed to the panoply of virulence
factors that are produced by them. The virulence factors can be broadly categorized into
the adhesive factors and toxins. They help UPEC to anchor in the urinary tract, invade the
urothelium and establish infective foci for persistence and resurgence in the organs of the
urinary tract.
III. Virulence Factors Produced by UPEC
Adhesins: UPEC produce a diverse range of adhesins such as P and type 1 pili
along with other fimbriae such as F1C, S, M and Dr (48). Out of these, the roles of P and
type 1 pili in the infection process have been studied in detail and hence are well
understood. Both P and type 1 pili are composite, multiprotein, hair-like appendages
3
emanating from the surface of UPEC. With their adhesive tips (PapG and FimH are the
adhesins for P and type 1 pili, respectively), they aid UPEC in colonizing the kidney and
bladder epithelia (7, 67). The PapG adhesin binds to α-D-galactopyranosyl-(α1-4)-β-Dgalactopyranoside moieties in the globoseries of kidney epithelium glycolipids while
FimH binds to manosylatyed glycoproteins in the uroplakins of the bladder epithelium
(69). Thus the bacteria get anchored in the urinary tract, an important advantage in face
of the gush of urine that frequently washes the urinary tract of any non-adherent particles.
Furthermore, FimH functions like a typical invasin by bringing about reorganization of
the actin cytoskeleton, localized alterations in the host cell membrane and uptake of the
bacteria (61).
The different variants of PapG, namely PapG-I, -II and -III bind to
different isoreceptors that vary in the number of N-acetylgalactosamine moieties added to
the Gal-(α1-4)-Gal core or by the addition of sialic acid residues (57). These variants are
proposed to aid in determining the host specificity of UPEC: PapGI and PapGII adhesins
preferentially bind globotriaosylceramide (GbO3) and globoside (GbO4), both of which
are common receptors on the human uroepithelium, respectively, whereas, PapGIII binds
Forssman Antigen (GbO5) receptor predominant on canine uroepithelium (57). P pili are
strongly associated with pyelonephritis, although reports describing the importance and
relevance of PapG adhesins and P pili in renal colonization and persistence are full of
inconsistencies: an isogenic Pap mutant in the pyelonephritogenic UPEC strain CFT073
caused acute pyelonephritis just as well as its WT counterpart in a CBA/J mouse model
of ascending UTI (66), whereas a PapG deletion mutant in another UPEC strain, DS-17,
failed to cause acute pyelonephritis in a primate (cynomolgous monkeys) model of
ascending UTI (80).
4
Toxins: Two important toxins that are coproduced and secreted by UPEC are
hemolysin A (hlyA) and cytotoxic and necrotizing factor 1 (CNF1).
CNF1 is a
dermonecrotic, 155 kDa, single chain AB toxin that increases the invasiveness of UPEC
by activating Rho GTPases (56). Active CNF1 is exported across the bacterial envelope
in a complex with outer membrane vesicles (17). It impairs remodeling and clustering of
CD11b receptors on the surface of polymorphonuclear leukocytes (PMNs), preventing
phagocytosis of opsonized UPEC (18). HlyA is secreted by the type 1 secretion system
(95). At high concentrations, HlyA targets erythrocytes and leukocytes to bring about cell
lysis (52), while at sublytic concentrations, it induces intracellular Ca++ signaling in
epithelial cells leading to pro-inflammatory responses such as production of IL-6 and IL8 (99). Secreted autotransporter toxin (Sat) is a cytolytic vacuolating toxin that has been
shown to damage the glomerular membrane and proximal tubule cells in the kidney, to
create lesions in intestinal and renal epithelial cell tight junctions and contribute to the
tissue damage associated with the UTIs (36, 37). In 2005, Aoki et al showed that UPEC
strain EC93 inhibits the growth of E. coli K12 strain MG-1655 in a mixed culture and
that the growth inhibition was dependent on the presence of surface protein contact
dependent inhibitor A (CdiA) and a two-partner secretion family member CdiB (4).
IV. Secretion Systems in Gram-Negative Bacteria
The virulence factors described above are produced in the bacterial cytoplasm.
For their assembly and/or secretion into the extracellular milieu, the virulence factors or
their components must be transported across the bacterial envelope with the help of
5
different secretion systems. The secretion pathways that are employed by Gram-negative
bacteria to transport cytoplasmic proteins across the bacterial envelope can be
categorized into two types. The Sec-independent type I, III and IV secretion systems
form a tunnel-like channel spanning the bacterial envelope such that the proteins are
exported from the cytoplasm into the extracellular environment in a one-step process
without the formation of periplasmic intermediates. On the other hand, Sec-dependent
secretion systems such as type II, chaperone usher and autotransporter pathways require
the Sec translocation machinery to transport cytoplasmic proteins across the IM into the
periplasm to form a periplasmic intermediate. The periplasmic intermediate is then taken
over by the secretion system for export across the OM (55). Figure 1.2 depicts different
secretion pathways in Gram-negative bacteria. Many of these pathways are utilized by
UPEC for the secretion of virulence factors. The chaperone usher pathway is utilized for
the assembly and secretion of adhesive pili on the surface of UPEC (96), while Sat is
secreted by means of the autotransporter pathway (37).
The type II secretion system (T2SS) is involved in the secretion of many
important toxic effectors in a diverse variety of Gram-negative pathogenic bacteria. The
protein components of the T2SS are similar to the structural and functional subunits of
the evolutionarily related type 4 pilus (T4P) system, which assembles adhesive
appendages on the bacterial surface that have been shown to aid bacterial colonization
and virulence. The main purpose of this thesis dissertation is to analyze the importance of
T2SS and T4P in the pathogenicity of the UPEC.
6
V. Type II Secretion System
In the 1980s, the T2SS was identified to be the mechanism involved in the transmembrane export of pullulanase in Klebsiella oxytoca (85). Type II secretion systems
(T2SS) are involved in the secretion of many important effectors in diverse Gramnegative pathogenic bacterial species (39, 47) that cause infections in humans, (cholera
toxin secretion mediated by Eps T2SS in Vibrio cholerae (86)), animals (aerolysin
secretion in the fish pathogen Aeromonas hydrophila) and plants (plant cell wall
degrading enzyme secretion in Erwinia caratovora) (84). In the case of enterotoxigenic
E. coli (ETEC), the T2SS secretes heat-labile enterotoxin (LT), which is structurally and
functionally similar to cholera toxin (94). E. coli K12 strains possess a T2SS that is
normally not expressed. However, in E. coli K12 mutants lacking the nucleoid structuring
protein H-NS, the T2SS was shown to be involved in the secretion of chitinase (28).
The T2SS is regarded as the main terminal branch of the general secretory
pathway (GSP). The T2S apparatus is a multiprotein complex that is coded for by a
cluster of 12-15 different genes (Fig. 1.3). In most species, the letters A through O and S
designate the T2SS genes and their protein products (see Fig. 3.1). The designations are
based on the K. oxytoca Pul system and the general identifier for each protein is Gsp (i.e.
GspA, GspB, etc.). The organization of the T2SS gene cluster is well conserved, with
most of the genes arranged to form a single operon (85).
The entire T2S apparatus, called the secreton (78), spans the bacterial envelope
(Fig. 1.3). It is interesting to note that even though the secreton is involved in the export
of bacterial proteins across the OM, most of its component proteins are associated with
the IM to form an IM platform of proteins (79). Structurally as well functionally the
7
subunits of a secreton are said to from three subassemblies: the OM pore formed by the
secretin (GspD), an IM platform of proteins, and the pseudopilus. GspD is a member of
the secretin family of gated pores that also includes proteins involved in T4P biogenesis,
filamentous phage extrusion, and the type III secretion system (85). The C terminus of
GspD is conserved across the secretin family of proteins and is embedded in the OM. On
the other hand, the N terminus is variable, periplasmic and possibly interacts with other
subunits of the secreton. An individual oligomer of GspD consists of 12-14 subunits.
GspS is a small OM lipoprotein subunit that promotes stabilization and OM insertion of
the secretin (GspD). GspE, a putative ATPase, is associated with the IM and provides
energy for opening the gated secretin or for assembly of the secretion apparatus (84).
GspG, H, I, J, and K exhibit homology with the structural subunit of type 4 pili, namely
the N-terminal region encoding the prepilin peptidase cleavage site. This subset of T2SS
proteins is called pseudopilins and they are N-terminally processed and methylated by the
T4P prepilin peptidase GspO. The pseudopilins of K. oxytoca T2SS (PulG-H) were
shown to assemble a pilus-like structure (pseudopilus) when the Pul secreton was overexpressed in E. coli K12 (89). Moreover, the high level expression of the endogenous
Gsp secreton of E. coli K12 brings about the assembly of GspG (a PulG homolog) into
bundled pili (100). Proteins GspF, L and M have been shown to interact with each other
to form a multi-protein complex at the IM that possibly acts as an anchor for the
pseudopilus (79).
The secreton is involved in the export of effector proteins across the OM only.
The effectors are transported across the IM by the Sec translocon or the Tat pathway
(101). In the periplasm, the effectors are completely folded. It is still unknown how a T2
8
secreton differentiates between its substrates and other periplasmic proteins. The secreted
proteins are structurally as well as functionally diverse. For example, elastase secreted by
P. aeruginosa is monomeric while cholera toxin is oligomeric; toxins like LT act inside
the host cells while hydrolytic enzymes degrade molecules such as lipids, chitin, and
pectin. The only defining characteristic of the T2SS effectors seems to be the relatively
high β sheet content in their fully folded form prior to secretion across the OM (85). The
failure to identify a linear signal sequence by comparing the primary amino acid
sequences of the effector proteins has led researchers in the field to believe that the signal
sequence is located at different parts of the linear protein and is brought together upon
protein folding. The search for a T2 signal sequence is full of contradictory findings. In
the case of PulA (pullulanase) secretion by K. oxytoca, the amino acid residues 1 to 78
(region A) and 734 to 815 (region B) were shown to retain ability to bring about the
secretion of β-lactamase (54) while the simultaneous deletion of these two regions
reduced the secretion of PulA to ≤ 25% (88). In a recent study, the PulA region between
amino acid residues 234 and 324 was shown to be important for its secretion (29).
Even though the actual process of secretion through the T2 secreton involves
participation by all three structural sub-assemblies; the exact role played by every single
component is not yet deciphered. In the current model for T2SS functioning, the GspE
ATPase is thought to provide the required energy for opening the OM gated pore GspD.
Energy from GspE is transmitted to the members of the IM platform of proteins which
span the periplasm contacting the GspD secretin and transporting the energy across the
periplasm (6, 77). The extension and retraction of the pseudopilus is believed to bring
9
about opening and closing of the OM secretin pore while the protein cargo is carried on
the tip of the pseudopilus (84).
VI. Type IV Pilus System
T4P are long, filamentous surface structures involved in myriad functions,
including social gliding and twitching motility, adhesion, biofilm formation, virulence
and horizontal genetic transfer by means of transformation, transduction and conjugation
(8, 74). Their polar localization, sequence conservation among the different pilins and
pilus assembly proteins, and role in bacterial motility sets them apart from other pili as a
different group (103). The most distinguishing feature of T4P is their ability to retract
through the bacterial envelope while their tips remain firmly adhered to a solid surface
acting as firm anchors and translocating the bacterial cell forward in a flagellaindependent manner (11, 64). T4P are known to play important roles in the virulence of a
variety of Gram-negative pathogens such as P. aeruginosa (62), Neisseria spp. (97) and
Vibrio cholerae (31). T4P of P. aeruginosa are a major virulence factor that aid bacterial
adherence to the tracheal epithelium (108). In N. gonorrhoeae as well as N. meningitidis,
association with the host epithelium followed by neisserial motility and dispersal over the
epithelial surface as well as induction of the cytoskeletal rearrangements at the site of
contact between Neisseria spp and host cells is mediated by T4P (63). Toxin-coregulated
type IV pili (TCP) of V. cholerae help the bacteria form microcolonies on the human
intestinal epithelium (42) and act as receptors for CTXΦ, the bacteriophage that carries
genes for cholera toxin (102).
10
T4P are extremely long (1-4 μm) and thin (50-80 Å) (15), but very strong
filaments that can withstand stress forces up to 100 pN (59). A type IV pilus fiber is a
homopolymer of pilin. The type IV pilin family of proteins consists of 18-22 kDa
polypeptides that are subdivided into two subgroups, A and B, based on whether the Nterminal amino acid in a mature protein is phenylalanine or methionine, respectively (92).
A pilin is synthesized as a precursor protein with a short (6-7 amino acid long; for group
A) or long (13-25 amino acid long; for group B) basic leader sequence that undergoes
processing by the preprotein leader peptidase/transmethylase leading to the production of
a mature pilin with an α-methylated phenylalanine or methionine residue at the Nterminus (72, 92, 93). Based on antibody mapping and X-ray crystallography studies, the
T4P fiber from P. aeruginosa and Neisseria spp. (type A) is proposed to form a 3-layered
structure: a central core of coiled conserved α-helices, surrounded by β-sheets, and the
outermost, most exposed hypervarible region of β-hairpins and extended C-terminal tails
(24).
The machinery involved in the biogenesis of T4P involves at least 16 protein
components closely related to those of the T2SS (Fig. 1.3). The protein PilB is the major
nucleotide binding ~62 KDa protein associated with the inner face of the cytoplasmic
membrane. Inactivation of pilB, or site-directed mutagenesis in the conserved Walker box
(consensus ATP-binding domain) domain ablates the assembly of T4P on the surface of
P. aeruginosa cells (98). PilT (as well as PilU in P. aeruginosa), a protein associated
with the cytoplasmic membrane is a PilB homolog and possesses the conserved Walker
Box. PilT is conserved among different organisms such as Myxocoxxus xanthus and N.
gonorrhoeae. Mutations in the ATP-binding domains of PilT (or PilU) result in
11
hyperfimbriated cells, which are unable to twitch (105, 106). This is reminiscent of
flagellar motor protein mutants that can assemble flagella which fail to generate
movement (103) and has given rise to a the idea of a bipartite motor in T4P with PilT
transducing energy for pilus retraction while PilB provides the energy source for pilus
extension by polymerization (103). PilQ belongs to the family of OM gated proteins
similar to protein D of T2SS and is involved in the final passage of polymerized PilA
pilins across the OM (60). Other protein components of the T4P biogenesis machinery
such as PilC, Pil M-P, PilZ and PilV, are known to be involved in T4P assembly and
secretion because mutants in any of these proteins are non-fimbriated, although the exact
function of each of these is unknown (1).
E. coli K12 possesses genes coding for all proteins required for T4P biosynthesis,
but under laboratory conditions the genes are transcribed at a very low level and pili are
not assembled. The major pilin subunit from E. coli K12, PpdD (prepilin peptidasedependent protein D) is assembled into pili if the Pul secreton is ectopically expressed
(89), or if ppdD is ectopically expressed in P. aeruginosa (87). ETEC employ T4P
(longus pili) as a colonization factor (34) while enteropathogenic E. coli (EPEC) use T4P
(bundle-forming pili) for localized adherence in the intestinal tract and autoaggregation
(33). Apart from the pseudopilins, the similarities between T2SS and T4P extend to
include other structural components, such as the OM secretin GspD, the ATPase GspE,
and GspF of the IM platform of proteins. In addition, a pilotin, GspS, is required for the
assembly as well as the stability of the OM secretin in both the systems (76). All these
observations indicate that T2SS and T4P are products of divergent evolution.
12
VII. Histology of the Urinary Tract
The major part of the urinary tract, from renal pelvis to the lumen of the bladder is
lined by the urothelium, which normally consists of three different cell layers: basal,
intermediate and superficial (81). The normal urothelium renews continuously but at a
very slow, 40-week turnover rate. The luminal layer of the urinary bladder is made of
highly differentiated, large and multinucleate facet or umbrella cells. Underlying this
layer are intermediate epithelial cells, sitting atop a layer of small, relatively
undifferentiated basal cells. A thin basement membrane and lamina propria separate the
urothelium from the connective tissue and serous layers of the outer wall of the bladder
(69). The apical face of facet cells is covered by a quasi-crystalline array of hexagonal
complexes of integral membrane proteins called uroplakins (9) to create a rigid-looking
asymmetric unit membrane (AUM) (81). So far five types of uroplakins (UPIa, UPIb,
UPII and UPIIIa and UPIIIb) have been identified. Their expression is differentiationrelated; i.e., the uroplakins are expressed in superficial and intermediate cells, but not in
the undifferentiated basal cells. The facet cell permeability barrier is supplemented with a
chemical barrier in form of the proteoglycans/ glycosaminoglycan that is known to
interfere with bacterial adherence (75). Towards the bladder neck, there is a sharp
transition from urothelium to truly stratified epithelium, which continues from this point
into the distal urethra and the point of urethral exit out of the body.
The renal papilla (inner medulla) is the part of the kidney where most of the water
from the glomerular filtrate is reabsorbed to form urine, which is delivered into the
ureters. To facilitate the process of water resorption, the inner medulla is lined by a
simple cuboidal epithelium that is one cell deep and devoid of uroplakins. The urothelium
13
extending into the other parts of the kidney such as outer medulla and cortex is 3-4 cells
deep and shows the presence of uroplakins in the superficial cells. A similar, 3-4 cell
deep, epithelial organization extends into the ureters with the umbrella cells and the
uroplakins presented on their apical surfaces lining the ureteral lumen.
The urothelium lining of the urinary tract is adapted to withstand the frequent
flushing by the urine. Apart from being a formidable permeability barrier that prevents
unwanted/toxic constituents of urine from leaching into host tissue and circulation (9),
the uroplakins also aid in adapting the bladder to the cycles of contractions and
distensions (81). This important structural defense is subverted by UPEC to cause
infections of the urinary tract.
VIII. UPEC Developmental Cascade and Urinary Tract Pathogenesis
The first important step in establishing a successful UTI is the attachment of
UPEC to the urothelium. The adherent bacteria are able to escape the cleansing action of
the gush of urine. Following adhesion, the UPEC can invade the superficial cells of
urothelium and form intracellular biofilm-like structures called pods, intracellular
bacterial communities (IBCs), or bacterial factories (68). The IBCs undergo a
developmental cascade that involves bacteria going through different morphological
changes (Fig. 1.5A) (50). Finally, the bacteria detach from the IBCs and flux out of the
infected cells. The formation and development of IBCs has been studied in great detail in
the bladder infection model and hence it is used as a paradigm for intracellular events
leading to UTI.
14
FimH-mediated Adhesion and Invasion: Isogenic UPEC mutants lacking the
FimH adhesin are defective in adhering to and invading the bladder in a mouse model of
cystitis as well as in cell culture using the human bladder carcinoma cell line, 5637, in a
gentamicin protection assay (61). 5637 cells also readily internalize FimH coated
polystyrene beads (61). Based on these findings, FimH adhesin present at the distal tip of
type 1 pili is considered to be important and sufficient for attachment of UPEC to the
superficial umbrella cells of the bladder epithelium. Uroplakin, UPIa, binds to the FimH
adhesin (69). Furthermore, FimH functions like a typical invasin by interacting with α3β1
integrins on the surface of the urothelial cells (19), followed by FimH-mediated
reorganization of actin cytoskeleton and localized alterations in the host cell membrane
(61).
Cytochalasin D, an inhibitor of F-actin polymerization, as well as specific
inhibitors of host protein tyrosine kinase, FAK (focal adhesion kinase) and
phosphoinositide-3 kinase (PI-3K) reduce FimH-mediated invasion. This has led
researchers to believe that FimH induces a pre-existing signal transduction cascade
involving phosphorylation of FAK followed by formation of FAK and PI-3K complexes.
This stimulates the generation of 3-phosphoinositide secondary messengers leading to
actin polymerization and cytoskeletal rearrangement. This in turn leads to bacterial
internalization (61).
In the superficial umbrella cells, UPEC spill out of the endosomes into the
cytoplasm and start replicating to form small loosely packed clusters called “early IBCs”
(50). The bacteria in early IBCs are rod-shaped (~3 μm in length), and divide rapidly
(doubling time of 30-35 min) for the first 8 hours after inoculation (50). On the other
15
hand, UPEC invading the underlying, less differentiated intermediate cells are trafficked
into vesicles that have properties of late endosomes or even lysosomes. An actin mesh
surrounds these vesicles and may impede bacterial growth (20).
Middle IBCs: From 8-10 hours after inoculation, the UPEC undergo many
changes; their growth rate drops, with doubling times of >60 min, the bacteria become
coccoid, with their average cell length of ~0.7 μm, and the IBCs become more tightly
packed (50). The growing IBCs push against the epithelial cell membranes giving it a
pod-like appearance as shown in Fig. 1.5B (3). Immunofluorescent staining for type 1 pili
and antigen 43, an autotransporter protein promoting agglutination, has shown a
heterogeneous expression of these two factors in pods. This has given rise to the idea of
IBCs being intracellular biofilms since the biofilms display community behavior with
different tasks delegated to different subpopulations of cells. Pods also show a strong and
uniform expression of polysaccharides. Presence of a polysaccharide matrix surrounding
the bacterial community is reminiscent of biofilms (3).
Late IBCs: As early as 12 hours post infection the bacteria on the outer edge of
IBCs differentiate into prototypical rod shape (average cell length 2 μm), become highly
motile, and dissociate from their parent IBCs (50). The motile bacteria have been
observed inside the umbrella cells as well as in the lumen of the bladder. Videomicroscopic evidence suggests that the intracellular motility is flagella-based. To begin
with, a few motile bacteria exit umbrella cells from a specific area on the cell surface. As
IBCs grow in size, the infected cells become apoptotic (53). The reduction in membrane
integrity of apoptotic cells might aid the motile bacteria to flux out. At some point the
16
entire IBCs burst out of the umbrella cells, releasing an estimated 105 bacteria into the
lumen of the bladder (50).
Filamentation of Bacteria: At around 20 hours post inoculation (PI), a
subpopulation of fluxed bacteria continues to grow without undergoing septation, giving
rise to long (up to 70 μm) and filamentous bacteria (50). The exact signal for
filamentation is unknown and some filamentous bacteria have been observed on the edge
of late IBCs along with the motile, detaching cells (50).
Advantages of IBC Formation in Uropathogenesis: Formation of intracellular
biofilm-like structures (IBCs) may provide the UPEC a safe haven away from the robust
immune onslaught that ensues in the wake of bacterial infection. The IBC developmental
cascade allows UPEC to establish replicating colonies in the infected bladders. Once the
IBCs mature, bacteria flux out in such large numbers that PMNs fail to clear the entire
bacterial population. Some of these bacteria are able to spread the infection.
Filamentation also aids the survival of bacteria that are fluxing out of infected cells
because PMNs do not respond to filamentous bacteria until the initiation of septation and
PMNs are unable to engulf the filamentous bacteria (50). However, filamentation alone is
not enough to subvert phagocytosis by PMNs, since bacterial filaments induced by
overproduction of a cell division inhibitor, SulA, are efficiently engulfed by PMNs (51).
IX. Recurrent UTI (rUTI)
More than 25% of UTI recur within 6 months (26). Their recurrence is attributed
to the ability of UPEC to inhabit underlying epithelial cells and to establish quiescent
17
reservoirs, from which bacteria can resurface at a later time point. In a murine model,
upon interacting with an intact epithelial barrier, UPEC were shown to form rosette-like
clusters inside early to late endosomes (Lamp1+/Rab5a-), called quiescent intracellular
reservoirs (QIRs), located within the umbrella cells. The induction of facet cell
exfoliation as part of host defense or as a result of the treatment with the cationic protein
protamine sulphate (PS) led to complete removal of infected urothelial cells harboring the
QIRs (70). Thus, at the end of 2 weeks post infection there was no evidence of bacteriuria
(presence of bacteria in the urine) and pyuria, the two important features of an active
infection. Concomitant with this, the superficial umbrella cell layer renewal process was
also complete. On the other hand, if the bladder epithelial barrier was partially damaged
(by administering PS) prior to the infection, it exposed the underlying transitional layers
to bacterial onslaught and QIRs were established inside the subepithelial layers of cells
(70). An actin mesh surrounds and impedes growth of UPEC within these vesicles (20).
Upon differentiation of the underlying transitional epithelial cells, the composition of the
actin mesh alters, leading to UPEC replication and eventually to the recurrence of UTI.
X. Host Response to Infection by UPEC
Introduction of UPEC into the bladder of a murine model of cystitis (C3H/HeN)
induces a robust inflammatory response. The Toll-like receptor (TLR) family plays an
important role in identifying LPS and other bacterial products. Different TLRs are
expressed on cells of the immune system such as macrophages, as well as the epithelial
cells. In mice lacking a functional tlr4 gene (C3H/HeJ), the IBC developmental cascade
18
progresses normally but without any appearance of neutrophils at the site of infection
(91). Also, UPEC emerging from infected umbrella cells of C3H/HeJ mice fail to form
filaments 48 hours after infection (50). The inability of C3H/HeJ mice that lack a
functional TLR-4 to mount an inflammatory response upon intravesicular challenge with
E. coli suggests that TLR-4 is the primary TLR that generates an inflammatory response
in the bladder (40). In 2004, TLR-11 was identified as a specific receptor for UPEC in
mice (107). But a bacterial ligand for TLR-11 is yet to be identified and it is still
unknown whether there is a human homologue of this receptor.
The cytoplasmic domains of TLRs (TIR) initiate a signaling cascade that ends in
the activation of NF-κB. Activated NF-κB goes to the cell nucleus and induces
transcription of anti-apoptotic and pro-inflammatory cytokines such as IL-6 and IL-8.
The presence of IL-6 and IL-8 in the urine of patients suffering from cystitis and
pyelonephritis indicates a possible role for these cytokines in UTI pathogenesis (69). The
epithelial cells of the urinary tract are a major source of these cytokines found in the
urine. Type 1 and P piliated UPEC have been shown to elicit stronger cytokine
stimulation as compared to their isogenic mutants. IL-6 is a pro-inflammatory or an
immuno-modulatory cytokine, which is known to facilitate transition from neutrophilic to
a predominantly monocytic response in many infectious diseases, although its role in UTI
is still unclear. IL-8 induction correlates with the appearance of PMNs in the urine
(pyuria), which is a hallmark of UTI. In mice, it has been shown that induction of MIP-2
(a murine homologue of IL-8) correlates with pyuria, underlining a role of IL-8 as a
neutrophil chemoattractant (41).
19
Many Gram-negative pathogens such as Yersinia and Salmonella spp. use the type
III secretion system to deliver effector molecules to interrupt the NF-κB signal
transduction cascade of the host (71, 73). It has been suggested that the ability of UPEC
to suppress NF-κB signaling might be advantageous in establishing a pathogenic cascade
in the urinary tract (especially in the bladder) despite a robust immune response. The
mechanics of such suppression must be different from that of Yersinia and Salmonella
spp. because UPEC lack the type III secretion system. The laboratory strain of E coli K12
(MG1655) activates IL-6 production in 5637 and T24 bladder epithelial cell lines in a
dose dependent manner (45). In contrast, IL-6 production when these cell lines are treated
with the UPEC clinical isolates UTI89 and NU14 is similar to unstimulated levels. These
UPEC strains also abolish the induction of IL-6 even when the cell lines (5637 and T24)
are treated with LPS or laboratory strains of E. coli (45).
XI. Model for UPEC Infections in the Bladder
A model depicting the various stages of UPEC infection in the bladder is shown
in Fig. 1.4. The FimH tip adhesin of type 1 pili on the surface of UPEC mediates the
attachment of bacteria to the uroplakins on the facet cells of bladder epithelium. The
soluble and cell associated factors in the bladder such as Tamm-Horsfall proteins,
secretory IgA, low molecular weight sugars and uromucoid act as anti-adhesive innate
immunity mediators. The FimH adhesin together with LPS also activates the host signal
transduction cascade involving protein tyrosine kinase, PI-3K and localized host actin
cytoskeletal rearrangements. This leads to the AUM of facet cells zippering around the
20
bacteria to bring about their internalization. The inside of a facet cell provides a safe
haven for the UPEC, away from the immune onslaught of the host. By an unknown
mechanism, the internalized bacteria escape into the cytoplasm, replicate and establish
the IBCs. In response to bacterial invasion, the facet cells, which otherwise have a very
slow turnover rate, exfoliate and get washed off with the flow of the urine. To avoid
being washed off with the exfoliating cells, the bacteria flux out of the infected cells and
invade the underlying layers of the bladder epithelium. The ability to move in and out of
the host urothelium can aid UPEC in disseminating to the other parts of the urinary tract.
The bacteria emerging out of the facet cells are often filamentous in form.
The
exfoliation of the superficial facet cells is thought to expose lower levels of the
urothelium to the bacterial infection. UPEC might persist in these cells for long periods,
in a quiescent state, to reactivate at a late time point causing a recurrent UTI.
XII. The Model Organisms
This project uses two clinical isolates of UPEC, CFT073 and UTI89, for the
diversity of clinical conditions they are known to cause. CFT073 was isolated from the
blood of a woman suffering from acute pyelonephritis and this strain has very high
cytotoxic and hemolytic activity (65). In contrast, UTI89 was isolated from a cystitis
patient (68). The sequencing of both of these organisms has been completed (12, 104).
Based on analysis of the genomic sequence, it is known that both CFT073 as well as
UTI89 contain a T2SS (similar to the K12 T2SS), along with genes coding for
components of T4P biosynthesis (12, 104). UTI89, but not CFT073, also possesses a
21
second T2SS gene cluster that is homologous to a T2SS present in enterotoxigenic E. coli
(ETEC) strain H10407.
Stanley Falkow formulated molecular Koch’s postulates in 1988 as guidelines to
characterize whether a specific gene is an essential component of a microorganism’s
ability to infect and cause disease in a particular host (23). These postulates state that the
specific inactivation of the gene(s) associated with the suspected virulence trait should
lead to measurable loss in pathogenicity and that the re-introduction of WT genes should
be accompanied by the restoration of virulence (22). Using the molecular Koch’s
postulates as guidelines, we generated deletion mutants in the essential T2SS and T4P
genes, compared them with their parental WT strains using different in vitro and in vivo
experiments, and complemented the defective phenotypes in mutants by introducing
plasmid-borne copies of the deleted gene. In lower organisms such as bacteria, the C
value (size of the haploid genome) (35) is proportional to the metabolic and
morphological complexity of the organism (10). In other words, one can argue that a
gene, let alone an operon, will not be present in the genome of a bacterium unless it has
some functional importance. Thus, in case of CFT073 and UTI89, what remains to be
found is the functions of the type II secretion and type IV pilus systems and their
relevance to the disease process.
22
FIG 1.1. Human urinary tract. A 3-D space filled rendering of human urinary tract
shows kidneys, which are the sites of blood filtration and urine formation; ureters, tubular
structures that transport the urine to the urinary bladder, a muscular sac that stores the
urine for excretion. (http://human-anatomy.net/urinary-tract-pictures.html)
23
Sec-independent Secretion
Sec-dependent Secretion
FIG. 1.2. Schematic representation of sec-independent and sec-dependent pathways.
In Gram-negative bacteria, virulence effectors produced in the cytoplasm have to pass
through the inner membrane (IM), periplasm and outer membrane to reach the
extracellular space and interact with host cells. This is accomplished by a variety of
secretion mechanisms. Well-known effector proteins secreted by these pathways are
shown. Hemolysin (HlyB) is secreted by the type I pathway in pathogenic E. coli, YopE
by type III in Yersinia pestis; VirE2 by type IV in Agrobacterium tumefaciens, P pili by
chaperone-usher pathway in UPEC; Immunoglobin1A protease by autotransporter in
Neisseria gonorrohoeae, and pullulanase (PulA) by type II in Klebsiella oxytoca. Figure
adapted from reference (55).
24
FIG. 1.3. Schematic representation of type II secretion and type 4 pilus biogenesis.
The components of type II secreton and T4P proteins are indicated using Gsp and Pil
nomenclatures, respectively. (a) T2SS effectors are transported across the IM in a Secdependent manner followed by the cleavage of the signal sequence, protein folding in the
periplasm, and the transport of the folded periplasmic intermediate across the OM
intermediate through the gated GspD secretin. GspC may transmit the energy generated
by the nucleotide binding protein, GspE, across the periplasm. (b) GspO or PilD is the
prepilin peptidase that is required for the processing of T2 pseudopilins (GspG-H) and
T4P pilin (PilA) and pilin like components such as PilE, PilV-X. (c) T4P fibers are
secreted across the OM through the PilQ secretin. The nucleotide binding proteins, PilT
and PilU provide energy for twitching motility while PilB energizes the pilus assembly.
Figure taken from reference (96).
25
FIG. 1.4. Progression of bladder infection by UPEC. UPEC expressing surface
adhesive pili adhere to uroplakins on the luminal surface of the superficial umbrella cells
aiding the bacteria to avoid getting purged by the flow of urine. Various host factors such
as Tamm-Horsfall protein competitively inhibit the interaction of the pilus adhesin with
uroplakins. UPEC are internalized in a FimH dependent manner, and replicate to form
IBCs. Apoptosis is induced and the infected umbrella cells are exfoliated. Meanwhile
IBCs mature and filamentous, motile bacteria emerge and disseminate infection to other
parts of the urinary tract. Figure taken from reference (69).
26
A
B
FIG. 1.5. UPEC developmental cascade in the host bladder. (A) Stages of IBC
formation and maturation are shown in a figure taken from a review by Kau et al (53). 1)
UPEC attachment to the uroplakins on the surface of facet cells (upper panel), and
eventual zippering of epithelial cells around UPEC is followed by 2) the formation of
IBCs as shown by hematoxylin and eosin (H&E) staining (upper panel). 3) During the
mid-IBC stage, UPEC interacts intimately with its matrix (upper panel) and also forms
luminal protrusions or pods (lower panel). 4) Late IBC stage is marked by the
filamentation of UPEC (upper panel) which later burst out of IBCs (lower panel). (B)
Scanning electron micrograph of pods formed on the surface of a mouse bladder. Scale
bar is 50 μm. Figure taken from Anderson et al (2).
27
Chapter Two
Materials and Methods
Bacterial Strains and Cell Lines
UPEC strain CFT073 was isolated from the blood of a woman suffering from
pyelonephritis (65), UTI89 from a cystitis patient (68), while MG1655 is a wellcharacterized laboratory strain of E. coli (5). The strains from the panel of pathogenic and
non-pathogenic E. coli isolates are listed in Table 2.1. The bacteria were grown and
maintained on LB agar plates or LB broth. When necessary, the medium was
supplemented with ampicillin (100 μg/ml), kanamycin (50 μg/ml), chloramphenicol (25
μg/ml) or arabinose (0.02%). Human bladder (5637 or ATCC#HTB-9) and human kidney
(A498 or ATCC#HTB-14) cell lines were obtained from ATCC. 5637 cells were
maintained in RPMI supplemented with 10% fetal bovine serum (FBS; Gemini) while
A498 cells were maintained in modified Eagle’s medium (MEM) supplemented with
10% FBS, 1X sodium pyruvate and 1X L-glutamine. For growth as well as during the
steps of infection assays the cell lines were kept at 370C in presence of 5% CO2.
Growth Curve Experiments
Bacterial cultures were grown overnight at 370C, 200 rpm in LB broth and diluted
in LB broth or tissue culture medium (RPMI +10% FBS) to OD600 < 0.1. The cultures
were maintained at 370C, 200 rpm for growth curve experiment in LB broth, or statically
at 370C in presence of 5% CO2 for tissue culture medium. The optical density of the
28
culture was measured at 600 nm using a spectrophotometer at indicated time points till
reading reached a plateau near the value of 1.
Enumeration of Bacteria
Infected cell lines or organs were treated with 0.4 and 0.02% triton X-100
(TX100) detergent in PBS, respectively. The cell monolayers were then kept at 40C for
10 min to help lift the cells off plastic while the organs were ground in a tissue
homogenizer. To obtain bacterial counts, the suspensions were vortexed, serially diluted
in sterile LB broth (100 μl of bacterial suspension was added into 900 μl sterile LB) and
100 μl of appropriate dilutions were spread on LB agar plates. After incubating the plates
at 370C overnight, single colonies were counted and CFU/ml estimated.
RNA Isolation and Semi-Quantitative Reverse Transcription (RT)-PCR.
Overnight cultures grown in LB were diluted 1:40 into LB or tissue culture
medium (RPMI + 10% FBS) and grown to early-to-mid log phase at 37ºC. LB cultures
were grown with aeration and bacteria in tissue culture medium were grown statically in
5% CO2. Bacteria were harvested after mixing with twice the volume of RNAprotect
reagent (Qiagen). RNA was isolated using RNeasy kit (Qiagen) as per manufacturer’s
instructions. Contaminating DNA was removed by on-column DNase treatment,
performed twice, using RNase-free DNase (Qiagen). RNA was eluted in RNase-free
water and the integrity of the RNA sample was checked by agarose gel electrophoresis
and ethidium bromide staining.
29
14 μl of DNase treated RNA was then mixed with 1 μl random primers
(Invitrogen), the mixture was kept at 950C for 2 min and then cooled to 420C over a
period of 20 min. 15 μl of mastermix containing 5X AMV reverse transcriptase buffer
(Roche), 0.3mM dNTPs and 50 U of AMV reverse transcriptase (Roche) was then added
and the mixture kept at 420C for 150 min. The cDNA was PCR amplified using specific
primers listed in Table 1. The PCR products were separated by agarose gel
electrophoresis, stained with ethidium bromide, and visualized using a Gel Doc 2000 gel
documentation system (Bio-Rad).
Generation of KO mutants
The method described in (16) was used to generate genomic deletion mutants
lacking the genes coding for OM secretins. This method harnesses the ability of the
bacteriophage λ recombination system (γ, β and exo) to bring about homologous
recombination as described in Fig. 2.2. Primers P0 and P1 were constructed such that
they contained, at their 3’ ends, 50 nucleotide long overhangs (H1 and H2)
complimentary to regions at the two ends of the target gene and are able to amplify a
region containing the antibiotic resistance cassette (ARC) carried on a template plasmid.
The PCR products were gel purified and cleaned to electroporation grade (Step I). In
parallel, a plasmid (pKD46, Red helper plasmid), carrying bacteriophage λ recombination
genes under an arabinose promoter, was electroporated into the strain in which the KO
was to be carried out. Replication of this plasmid is temperature-sensitive. In the presence
of arabinose, enzymes necessary for recombination were ectopically produced inside the
cell. These bacteria were then electroporated with the PCR products carrying the ARC.
30
Homologous recombination and concomitant replacement of the gene by the ARC was
determined by the ability of bacteria to grow on medium containing antibiotic (Step II).
The bacteria were then maintained in a medium lacking antibiotic at a temperature that
does not allow pKD46 replication. Subsequently, drug resistant bacteria were made
electro-competent and electroporated with a plasmid (pCP20) carrying the FLP gene
under the control of a natural promoter (Step III). At the permissive temperature, the FLP
recombinase was induced to eliminate the ARC. Cells were cured of pCP20 by growing
them at a temperature non-permissive for its replication (Step IV).
Adhesion, Invasion and Efflux Assay
To cause a successful UTI, UPEC have to adhere to the host urothelium, invade it,
form intracellular replication foci, escape the facet cells that are being shed as part of
innate immune response and infect underlying layers of cells. These four cardinal steps in
UPEC mediated UTI, namely adhesion, invasion, intracellular growth and efflux
(extrusion) are recapitulated in a modification of the gentamicin protection assay (61) the
procedure for which is described below. We used this assay protocol to compare different
KO mutants with their WT counterparts in order to look for any differences in one or
more of the four characteristic steps in UTI.
Human bladder or human kidney cell line monolayers in 24 well plates were
infected with the appropriate bacteria grown in static cultures at 370C for 48 hours. Prior
to infection, the number of bacteria was normalized to OD600~0.5 and enumerated
further by plating the serial dilutions of bacteria on LB agar plates; initial count (IC). 10
μl/ml of the normalized bacteria were used for infection to achieve an MOI of 5-8. The
31
bacteria were spun down at 800 X g for 5 min to expedite their contact with host cell
monolayers. The infected monolayers were kept at 370C in presence of 5% CO2 for 2 h.
For the adhesion-invasion module of this assay, the monolayers from one set of
wells were washed five times with sterile PBS containing CaCO3 and MgCO3
(Ca++/Mg++), lysed with 40 μl of 10% TX100. The intra-cellular as well as adherent
bacteria in these lysates were enumerated; adherent bacteria (AB). The second set of
wells that was left unwashed was lysed with TX100. The intracellular, adherent as well as
non-adherent suspended bacteria in the lysates were enumerated; total bacteria (TB).
Adherence frequencies were calculated as %AB ÷ TB.
The third set of wells was treated with 100 μg/ml gentamicin for 2 hours at 370C
in the presence of CO2. Further, this was washed five times with PBS (Ca++/Mg++),
treated with TX100 and intracellular bacteria (IB) were enumerated. Invasion
frequencies were calculated as %IB ÷ IC.
For the efflux assay module, the infection steps were carried out just as in the
adhesion-invasion module except that after the 2-hour treatment with 100 μg/ml
gentamicin, the infected monolayers were washed with PBS (Ca++/Mg++) once and were
maintained in presence of 10 μg/ml gentamicin overnight at 370C in 5% CO2. On day 2,
the intracellular bacteria in one set of wells were enumerated after washing off
gentamicin and lysing the cells with TX100; overnight intracellular bacteria (OB). The
other two sets of wells were washed 5 times with PBS (Ca++/Mg++) and kept for 6 hours
at 370C, 5% CO2 either in the presence or absence of gentamicin. In the absence of
gentamicin, the intracellular bacteria form the IBCs and start fluxing out of the infected
cells. These cells were treated with TX100, and the suspensions plated to obtain the total
32
number of intracellular and effluxed bacteria (EB). Efflux index was calculated as %EB
÷ OB. The intracellular bacterial counts (IC-6) at the end of the 6 h incubation period
were determined from the wells containing gentamicin. The intracellular growth index
over a period of 6 h was determined as % IC-6 ÷ OB.
For an extended intracellular growth experiment, the infected steps were carried
out as in the adhesion-invasion module, followed by washing with PBS (Ca++/Mg++) and
maintaining the infected monolayers in presence of 10 μg/ml of gentamicin. The
intracellular bacteria were enumerated after washing the monolayers and lysing the cells
using TX100, at indicated time points.
Mouse Infections
Static cultures of bacterial strains were grown in 20 ml LB broth at 370C for 48
hours. On the day of animal infection, the bacteria were centrifuged and pellets
resuspended in sterile PBS.
6-8 weeks old CBA/J female mice were infected with 50 μl of bacterial
suspensions normalized to ~109 CFU/ml. The mice were anesthetized using isoflurane
and the 50 μl bacterial suspension or sterile PBS (as negative control) was delivered to
the bladder by transurethral catheterization using UV sterilized 0.28 mm inner diameter
tubing (Becton Dickinson cat # 427401) mounted on 30 ½G needle tips. Dilutions of the
inocula were plated on LB agar to determine input CFU/ml for all bacterial strains. 7
mice were infected with each bacterial strain. On day 7 PI, the animals were killed by
CO2 asphyxiation, dissected and organs (bladders and kidneys) harvested. Bladders and
33
right side kidneys from 5 infected mice were homogenized in sterile PBS containing
0.02% TX100, and serial dilutions were plated on LB agar plates.
Bladders and right side kidneys from the other two infected mice were stored in
paraformaldehyde for tissue sectioning. Sections were stained by hematoxylin and eosin
(H&E). The left side kidneys were homogenized, serial-diluted and CFUs enumerated.
Each mouse experiment with the UTI89 and CFT073 knockout mutants was repeated at
least once. Comparison of the UTI89 WT, ∆hofQ, and ∆hofQ/phofQ strains was
performed once. Comparison of the CFT073 WT, ∆yheF ∆hofQ, ∆yheF ∆hofQ/pyheF
and ∆yheF ∆hofQ/phofQ strains was performed twice.
All animal work was conducted according to the guidelines of the Stony Brook
University institutional animal care and use committee (IACUC# 2005-1465).
Generation of Rescue Plasmids
The genes hofQ and yheF were PCR amplified using cloning primers listed in
Table 2.2. The PCR products were gel purified, ligated into pGEM® T-easy vector
(Promega) as per the manufacturer’s instructions and the ligation products were
transformed into DH5α. The transformants containing correct ligation products were
identified by their ability to grow in the presence of 100 μg/ml ampicillin. The colonies
containing pGEM-hofQ and pGEM-yheF were purified, the plasmids isolated and
analyzed by DNA sequencing. EcoRI digested fragment from pGEM-hofQ and BamHI,
HindIII digested fragment from pGEM-yheF were ligated into pMMB67HE pretreated
with the same restriction endonucleases to obtain plasmids, phofQ and pyheF. The correct
orientation of hofQ gene in pMMB67 was confirmed by the DNA sequencing. These
34
plasmids were then transformed into appropriate deletion mutant strains for use in the
rescue experiments.
IL-6 Suppression Assay
UPEC suppress the induction of IL-6, a pro-inflammatory or immunomodulatory
cytokine, produced by the urothelium in response to Gram-negative bacterial LPS,
laboratory strains of E. coli and IL-1 (45). We compared different KO mutants with their
WT counterparts for ability to suppress IL-6 production by the human bladder cell line,
5637 using the method described by Hunstad et al (45).
Static cultures of bacteria were grown overnight at 370C in 20 ml LB broth. 48
hour prior to the experiment, 5637 bladder cells were treated with trypsin-EDTA, washed
with tissue culture medium (RPMI+10%FBS) and seeded into 24-well plates to get
confluent monolayers. On the day of experiment, bacteria were centrifuged, the pellet
washed in pre-warmed sterile PBS, and resuspended in PBS containing such that there
were 109cfu/ml (bacteria were enumerated by plating serial dilution on LB agar plates to
determine the numbers of live bacteria added to each well). Bladder cell monolayers were
washed with sterile PBS once and fresh tissue culture medium containing 5 μg/ml O157
LPS (Sigma) applied. 10μl normalized bacteria were added to each well. The bacterial
contact with cells was expedited by centrifuging the plates at 400 X g for 3 min and the
plates incubated at 370C in presence of 5% CO2 atmosphere for 12 hours. At the end of
incubation period, the supernates were extracted, spun at 13000 X g for 5 min to pellet
bacteria as well as cell debris and stored at –800C until IL-6 estimation by ELISA
(Antigenix America).
35
Cytotoxicity Assay
Cytotoxicity induced by different knockout mutants and their WT counterparts
was compared under two different conditions; during the period of efflux after
gentamicin removal and in the first three hours after infecting the 5637 human bladder
cell line.
To measure cytotoxicity during bacterial efflux, 5637 human bladder cells were
infected with WT or knockout UPEC strains for an efflux assay as described above. Upon
removal of gentamicin on the second day of the infection, the supernatant fractions were
collected at 2, 4 and 6 h time points. These fractions were tested for the presence of LDH
using the Cytotox 96 non-radioactive cytotoxicity assay (Promega). In separate
experiments, LDH release was determined after infecting 5637 cells for 3 h with different
amounts (104 to 106 CFU/ml) of WT or knockout UPEC strains.
Biofilm Assay
Overnight grown (200 rpm, 370C.) bacterial strains were diluted 1:100 and 100 μl
of diluted culture was added to a 96-well, U-bottom, polystyrene plate in quadruplicate.
P. aeruginosa was used as a positive control, medium only wells as negative control, and
3 wells were left empty as blanks. The plates were then incubated under stationary
growth conditions at 370C for 48 h. At the end of the incubation period, the wells were
washed with PBS three times, vigorously, to remove all planktonic bacteria. 150 μl of
0.1% crystal violet was added to all the wells except the blanks and incubated for 10 min
at room temperature to stain the biofilms. Later, the wells were washed three times with
water, until no color was left in the negative control wells. Plates were then inverted on a
36
paper towel for 30 min to dry the wells completely. To solubilize the stained biofilms,
200 μl solvent solution (80% ethanol and 20% acetone) was added to all the wells
including blanks, and the covered plates were incubated at RT, shaking for 15 min. 150
μl of the solubilized biofilms were transferred to fresh 96-well plates and the absorbance
measured at 575 nm. OD575 ≥ 0.8 is considered positive for biofilm formation.
Statistical Analysis.
The CFU/g of organ weight results from the mouse infection experiments were
compared using the Mann-Whitney test for non-parametric data with one-tailed P value.
The cell culture assay indices (efflux, adhesion, and invasion) were compared using
Dunnett’s multiple comparisons test. P values < 0.05 were considered significant.
37
Table 2.1 List of strains from the panel of pathogenic and non-pathogenic E. coli.
Clinical Isolate
SNAC-396
SNAC-403
VVC-117
C&C-1022
C&C-1033
Pylo-27
Pylo-28
Pylo-29
Pylo-30
Pylo-31
Pylo-33
Pylo-34
Pylo-36
Pylo-37
Pylo-38
Pylo-39
Pylo-41
Pylo-45
Pylo-49
Pylo-50
TOP-9
TOP-14
TOP-31
TOP-13
TOP-21
TOP-2
TOP-11
TOP-48
TOP-49
TOP-54
TOP-80
TOP-81
TOP-82
TOP-85
TOP-86
TOP-89
U01-85
U01-87
U01-89
U01-90
U01-92
U01-93
U01-101
Associated Pathological Condition
First UTI
First UTI
First UTI
First UTI
First UTI
Pyelonephritis
Pyelonephritis
Pyelonephritis
Pyelonephritis
Pyelonephritis
Pyelonephritis
Pyelonephritis
Pyelonephritis
Pyelonephritis
Pyelonephritis
Pyelonephritis
Pyelonephritis
Pyelonephritis
Pyelonephritis
Pyelonephritis
Asymptomatic Bacteriuria
Asymptomatic Bacteriuria
Asymptomatic Bacteriuria
Asymptomatic Bacteriuria
Asymptomatic Bacteriuria
Recurrent UTI
Recurrent UTI
Recurrent UTI
Recurrent UTI
Recurrent UTI
Rectal Isolate
Rectal Isolate
Rectal Isolate
Rectal Isolate
Rectal Isolate
Rectal Isolate
Cystitis
Cystitis
Cystitis
Cystitis
Cystitis
Cystitis
Cystitis
Sourcea
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
TMH
38
Clinical Isolate
U01-105
U01-106
U01-107
U01-108
EC-42
EC-93
EC-45
EC-55
EC-56
B-217
B-212
B-220
B-223
M-1
M-2
Associated Pathological Condition
Cystitis
Cystitis
Cystitis
Cystitis
Cystitis
Cystitis
Cystitis
Cystitis
Cystitis
Cystitis
Cystitis
Cystitis
Cystitis
Meningitis
Meningitis
Sourcea
TMH
TMH
TMH
TMH
MAM
MAM
MAM
MAM
MAM
MAM
MAM
MAM
MAM
Lab
Lab
a
The clinical isolates from the panel were supplied by Dr. Thomas M. Hooton (TMH),
University of Miami or by Dr. Matthew A. Mulvey (MAM), University of Utah.
39
Table 2.2 List of primers used in this study.
Primer
Sequence
Size of PCR
Product (bp)
WT
Knockout
T4P Secretin (hofQ)
Knockout
forward (P0a)
Knockout
reverse (P3a)
Flank A
Flank B
RT-PCR
forward
RT-PCR reverse
Clone forward
Clone reverse
TTACTCACTGGAAACCAGTCGTGGCGTG
ATAAACACCACTAACTCGCGTCGTGTAG
GCTGGAGCTGCTTC
ATGAAGCAATGGATAGCCGCACTACTGT
TGATGCTGATACCCGGCGTACACATATGA
ATATCCTCCTTAGT
CATATTGAGTTGTTGAGCTAACTGG
1724
GACAATTTTACAGCTGACGCC
726
GATGGTGGATGACGTTCCG
650
AACGTCCGTTGATGCGTCC
CAACAAGCTTGTGAGCCGTTTTTCGTTA
AGCGTGG (HindIII)
GATAAAGCTTCATCATGGAACGCAGGCA
GC (HindIII)
ETEC-T2SS secretin (gspD)
Knockout
forward (P0a)
Knockout
reverse (P2a)
Flank A
Flank B
RT-PCR
forward
RT-PCR reverse
TTAACGCGTTCTCCCGGCATTGAGGAAC
GCGCGTACTTCCGGCGGTAAGGTGTGTA
GGCTGGAGCTGCTT
GTGTTTTGGCGTGATATGACGTTGTCTAT
CTGGCGTAAGAAGACAACTGGCATATGA
ATATCCTCCTTA
AGAGAGCGACAACGGATGAACATGG
2343
CGCCATTGCGCTAAATCAGC
323
CATACCGTCACGCAAAATGGTC
899
CCAGCAAACACAGTAATGCCCTG
K12-T2SS secretin (yheF)
Knockout
forward (P0a)
Knockout
reverse (P2a)
Flank A
Flank B
RT-PCR
forward
RT-PCR reverse
GGACTCAATAAAATCACCTGCTGCTTGC
TGGCAGCACTACTCATGCCTTGTGTGTAG
GCTGGAGCTGCTT
TCACCGTGACGATGGCGCAGGAGCGTG
ACTGTTGAACGTGTTCTCATCCACATATG
AATATCCTCCTTA
GAACTTCAGGAATGAATGG
2018
TGAATTCTCATAAGAATGCC
194
TGCAGGACACGCTGAGAACG
350
GAACGTTCTCAAGCGGTACG
40
Primer
Sequence
Size of PCR
Product (bp)
WT
Knockout
Clone forward
TGCTAAGCTTACAATTCGTCGCAACGGC
(HindIII)
Clone reverse
AGCGGGATCCTGGCGGGGTACGGTGAG
TG (BamHI)
Type 1 pilus usher (FimD)
Knockout
forward (P0a)
Knockout
reverse (P2a)
Flank A
Flank B
ATGTCATATCTGAATTTAAGACTTTACC
AGCGAAACACACAATGCTTGCAGTGTAG
GCTGGAGCTGCTTC
TTAACGACATTCAGGCTGATAGCTGGGT
TAATAACTGCTGCTGACTCTCTGCATATG
AATATCCTCCTTAGT
CGCACTTACCCCCAAAATGAC
ATGGAACAACAGGAGTCGTCGC
2959
16S rRNA
Forward
TGACGGGGGCCCGCACAAGC
Reverse
CGGCCGGACCGCTGGCAACA
E. coli Panel (ETEC-T2SS specific primers)
J Forward
CAACCAGATGCGTTCAATTTCG
J Reverse
TCTGACACAAATGATGCCGC
D Forward
TTAAGCCTTGCTCATCGCGG
D Reverse
CTTCTTCATGGTGGGCCAGG
222
450
500
a
The primers P0, P1, P2and P3 are italicized.
The restriction sites introduced in the primers are indicated in parentheses and their
location within the primer sequence is underlined.
b
41
H1
Step I
P0
A n tib iotic R esistan ce G en e
P2
H2
λR ed
P A ra
H1
Step II
P2
P0
F RT
A
H2
30 0 C.
pK D46
F RT
B
Step III
K1
H1
H2
FLP
K2
30 0 C .
pCP 20
Step IV
H1
H2
FIG. 2.1. Schematic representation of λRed recombinase-based gene deletion. P0 and
P2 are priming sites in the antibiotic resistance cassette carried on a template plasmid. H1
and H2 are 50-60 nucleotide long regions homologous to regions of the target gene. A
and B represent the flanking primer pair and K1 and K2 are ARC-specific primers. These
were used for PCR verification of different variants and junction fragments formed
during the procedure. Plasmid pKD46 carries bacteriophage λ recombination genes under
the control of an arabinose promoter while pCP20 carries the FLP recombinase gene
under the control of a natural promoter. Both of these plasmids replicate only at 300C
(16).
42
Chapter Three
T2S and T4P Systems Contribute to the Virulence of UPEC
I. INTRODUCTION
E. coli are a diverse group of Gram-negative bacteria that can be subdivided into
three major groups: commensals in the intestinal tract of warm-blooded mammals, agents
of intestinal infections such as hemorrhagic colitis and hemolytic-uremic syndrome, and
ExPEC (52). ExPEC are causative agents of a range of diseases, including UTI,
nosocomial pneumonia, neonatal meningitis, and neonatal sepsis. UTIs comprise a group
of disorders, including cystitis and pyelonephritis, which is defined by the presence of
microorganisms in the otherwise sterile urinary tract and UPEC is the principal causative
agent.
Many Gram-negative bacterial pathogens use type II secretion and the
evolutionarily related T4P system to secrete and/or assemble virulence factors. Although
both T2SS and T4P have well-studied roles in the virulence of intestinal pathogenic E.
coli (33, 34, 94), their roles in the virulence of ExPEC had not been investigated. In this
dissertation research, we investigated the roles of T2SS and T4P in the pathogenesis of
the model UPEC strains CFT073 and UTI89, isolated from patients suffering from
pyelonephritis and cystitis, respectively (65, 68). Both model strains, CFT073 and UTI89,
possess a T2SS along with genes coding for components of T4P biosynthesis, which are
similar to the K12 T2S and T4P systems (12, 104). Moreover, UTI89 possesses a second
T2SS gene cluster that is homologous to a T2SS present in ETEC strain H10407. We
43
generated knockout mutations in OM secretin genes of the UPEC T2S and T4P systems
and carried out direct comparison between wild type (WT) strains and knockout mutants
in vivo using the CBA/J mouse model of ascending UTI. Our findings identify T4P as
virulence factors of UTI89, and T4P and/or the T2SS as virulence determinants of
CFT073. In particular, our results show that these systems are important for persistent
infection of the upper urinary tract.
44
II. RESULTS
T2SS and T4P gene clusters in CFT073 and UTI89. UPEC strain CFT073 is a
well characterized strain isolated from the blood of a woman suffering from
pyelonephritis (65), whereas UTI89 is a UPEC strain isolated from a cystitis patient (68).
Annotated complete genome sequences for both these model clinical isolates have been
published (CFT073, accession NC_00431; UTI89, accession NC_007946) (12, 104). The
T2SS and T4P genes present in the UTI98 and CFT073 genomes are shown
schematically in Fig. 3.1. Both UPEC clinical isolates possess gene clusters homologous
to a T2SS gene cluster present in E. coli K12 (K12-T2SS). Both UPEC strains also
possess genes homologous to E. coli K12 T4P genes. The T2SS and T4P genes common
to both UTI89 and CFT073 have very high amino acid homology to the corresponding E.
coli K12 systems (91-99% identity, except for the gspI minor prepilin, which is 84%
identical). In addition to the K12-homologous gene clusters, UTI89 possesses a second
T2SS gene cluster with 84-97% amino acid homology to a cluster present in ETEC strain
H10407 (ETEC-T2SS). This second UTI89 T2SS has relatively low homology to the
K12-T2SS (16-60% amino acid identity). The ETEC-T2SS cluster is not present in strain
CFT073.
RT-PCR was used to ascertain expression of the UPEC T2SS and T4P genes
under various culture conditions, including growth to log phase in LB or tissue culture
medium (RPMI + 10% FBS). Expression of T2SS and T4P genes in the E. coli K12
strain MG1655 was used for comparison. Fig. 3.2 shows the expression level of
transcripts corresponding to the OM secretins. Both the yheF (K12-T2SS) and hofQ
(T4P) secretin genes were expressed at low levels in UTI89 and CFT073, with expression
45
higher in LB than in tissue culture medium. Expression of these genes in the UPEC
strains was comparable to expression detected for MG1655. Interestingly, the ETECT2SS secretin gspD was expressed at relatively high levels in both LB and RPMI,
suggesting this system is activated under these culture conditions.
Construction of T2SS and T4P OM secretin deletion mutants. To analyze the
importance of the T2SS and T4P genes in the pathogenesis of UPEC, we targeted the OM
secretin of each system for deletion mutagenesis using the bacteriophage λ Red
recombinase system (16). The secretin component forms a gated channel required for
export of T2SS effectors or T4P across the OM. We made single knockout mutations of
each of the three secretins present in UTI89, namely yheF for the K12-T2SS, gspD for
the ETEC-T2SS, and hofQ for T4P (Fig. 3.1). For CFT073, we constructed single
knockout mutations of the yheF (K12-T2SS) and hofQ (T4P) secretins, as well as a
double deletion strain lacking both secretins (Fig. 3.1). Proper construction of the
knockout mutations was verified by PCR (Fig. 3.3) using primers specific for the
flanking regions of the targeted genes (Table 2.2). We also checked each of the knockout
mutants for growth in LB, RPMI supplemented with 10% FBS (Fig. 3.4) and in M63
minimal medium (data not shown); growth of each mutant was similar to the parental WT
strain.
Roles of the UPEC T2SS and T4P in pathogenesis. To determine the roles of
the T2SS and T4P in the ability of UPEC to successfully infect the urinary tract, we
tested the UPEC secretin knockout strains using the CBA/J mouse model of ascending
UTI (38). For UTI89, we inoculated 107 CFU of the WT strain or one of the secretin
mutants into the bladder by trans-urethral catheterization. At days 1 and 7 PI, animals
46
were sacrificed and CFU/g of bladder or kidney were determined. All knockout mutants
colonized the bladder and kidneys to similar levels as WT UTI89 on day 1 PI (data not
shown). On the other hand, while all knockout mutants were able to colonize the bladder
similar to WT on day 7 PI (Fig. 3.5A), the ∆hofQ (T4P) mutant was significantly
defective in colonizing the kidneys on day 7 PI (p=0.03, Fig. 3.5B). The ∆yheF (K12T2SS) and ∆gspD (ETEC-T2SS) mutants colonized the kidneys similar to WT UTI89
(Fig. 3.5B). This defines a specific function for T4P in colonization of the kidney.
We next performed similar experiments with the CFT073 secretin mutants. 107
CFU of WT CFT073, the ∆yheF (K12-T2SS) or ∆hofQ (T4P) single secretin mutant, or
the double secretin mutant (∆yheF ∆hofQ) were inoculated into the bladder of CBA/J
mice by trans-urethral catheterization. As found for UTI89, none of the CFT073 mutants
showed defects in colonizing the bladder or kidneys compared to WT CFT073 on day 1
PI (data not shown). The single secretin knockout mutants of CFT073 also showed no
defects in colonizing the urinary tract on day 7 PI (Fig. 3.5 C and D). However, CFU/g
recovered from kidneys on day 7 PI of mice infected with the CFT073 ∆yheF ∆hofQ
double mutant were significantly lower (p=0.015) compared to WT CFT073 (Fig. 3.7D).
In addition, we noted consistently reduced, though not statistically significant, median
CFU/g recovered from the bladders of mice infected with the double knockout strain
(Fig. 3.5C). These results demonstrate a role for T4P and/or the T2SS in the ability of
CFT073 to persist within the urinary tract.
Next, we checked whether the renal persistence of UTI89 ΔhofQ and CFT073
ΔyheF ΔhofQ mutant strains could be restored back to WT-like levels by introduction of
the deleted genes in trans. The genes yheF and hofQ were cloned into the low-copy
47
number plasmid pMMB67HE as described in Materials and Methods. The resulting
plasmids pyheF and phofQ were used to complement the defective phenotype in UTI89
and CFT073. Complementation of UTI89 ΔhofQ with phofQ restored the colonization of
renal tissue in the CBA/J mouse infection model to WT-like levels (Fig 3.6A). To
confirm the specificity of the kidney phenotype for the CFT073 ∆yheF ∆hofQ double
mutant, we complemented this strain with either yheF or hofQ in trans (using plasmids
pyheF or phofQ, respectively). Complementation with phofQ restored kidney
colonization to WT-like levels (Fig. 3.6B). In contrast, although complementation with
pyheF increased bacterial colonization in the kidney, this did not reach statistical
significance (Fig. 3.6B). These results demonstrate a specific role for T4P in the ability of
CFT073 to persist within the urinary tract, in particular the kidneys, and suggest the K12T2SS may also contribute to kidney colonization.
Histological examination of the infected organs. To gain further insights into
the defective nature of UTI89 ΔhofQ and CFT073 ΔyheF ΔhofQ in terms of the reduced
levels of persistence in the kidneys, bladders and the right side kidneys from the animals
inoculated with these mutant strains, their WT counterparts, or sterile PBS were fixed in
formaldehyde, sectioned and stained with H&E. The CFUs from the corresponding leftside kidneys were enumerated to determine the extent of infection. As shown in Figs. 3.7
and 3.8, the kidney sections from mice infected with WT UPEC strains exhibited signs of
severe inflammation, especially of the renal pelvis (the central part of the kidney where
the urine accumulates prior to discharge), marked by the infiltration of papillary spaces
and collecting ducts by inflammatory cells such as neutrophils. In contrast, kidneys
obtained from mice inoculated with PBS exhibited normal tissue architecture as indicated
48
by clear papillary spaces, calyx, collecting ducts and the symmetrical nature of the
glomerulus. The histological examination of renal sections obtained from mice infected
with UTI89 ΔhofQ did not show any signs of inflammation (Fig. 3.7C). The kidney
sections obtained from a mouse infected with CFT073 ΔyheF ΔhofQ looked free of
inflammatory cells, similar to the PBS inoculated kidneys (Fig. 3.8C) and the left-side
kidney from this animal did not yield CFUs. In contrast, the kidney section from the
second CFT073 ΔyheF ΔhofQ infected mouse showed WT-like inflammation (Fig. 3.8D).
The corresponding left-side kidney from this mouse yielded high CFUs (1.2 X 105 CFU/g
of kidney). Thus, there was an obvious correlation between inflammation and the
presence of detectable pathogens in the renal tissue. Consistent with the lack of detectable
phenotype in the bladder, all bladder sections presented similar signs of inflammation,
whether the mice were infected with WT or mutant strains (data not shown).
49
III. DISCUSSION
Many Gram-negative bacterial pathogens use T2S and T4P systems to secrete
and/or assemble virulence factors. The two model clinical isolates CFT073 and UTI89
each possess gene clusters that have close homology to T2SS and T4P genes in E. coli
K12 while UTI89 contains a second T2SS homologous to one found in ETEC strain
H10407 (Fig. 3.1). To understand the importance of these systems in the virulence of
UPEC, we constructed deletion mutations in T2SS and/or T4P OM secretin genes. We
verified the specificity of the gene-deletion by carrying out DNA-PCR using different
flanking and insertion-specific primers (Fig. 3.3). The deletion mutants did not exhibit
any defects growing in LB, RPMI + 10% FBS or M63 minimal medium (Fig. 3.4). This
is important because a general growth defect in a deletion mutant would have prevented
its inclusion in the in vivo experiments.
We compared the secretin mutants to their WT counterparts using different
animal model infection experiments. In the CBA/J mouse model of ascending UTI, we
found that only the ΔhofQ (T4P) mutant of UTI89 was significantly defective in
colonization of kidneys on day 7 PI as compared to WT UTI89 (Fig. 3.5B), although
there was no apparent defect in the ability of this strain to colonize and persist within the
bladder (Fig. 3.5A). Thus, T4P were required for UTI89 to colonize and persist in
kidneys. The kidney colonization defect of the ΔhofQ mutant was rescued by
reintroduction of WT hofQ gene (Fig. 3.6A), further establishing a specific role for T4P
in the ability of UTI89 to persistently colonize the kidney. In the pyelonephritogenic
isolate CFT073, mutants containing single deletions of ΔhofQ or ΔyheF were able to
infect and colonize bladder and kidney tissues just as well as the WT strain. However,
50
similar to the UTI89 ∆hofQ strain, a CFT073 double knockout mutant lacking both yheF
and hofQ showed significantly reduced colonization of the kidney on day 7 PI (Fig.
3.5D). Thus, the two OM secretins of CFT073, HofQ and YheF, may be able to replace
one another functionally and hence the kidney defect was observed only when the mice
were infected with a double deletion mutant. This possibility is further supported by the
results of the complementation experiments, in which reintroduction of hofQ fully
restored kidney colonization, but expression of yheF in trans only partly rescued the
defect (Fig. 3.6B). Thus, T4P are likely the critical virulence factor for CFT073, as found
for UTI89, but YheF is able to compensate at least partially for loss of the HofQ pilus
secretin. Lack of full complementation by pyheF could be attributed to expression of
yheF from an artificial promoter on the plasmid. Further investigation is required to
determine the specific contributions of T4P and the T2SS to the pathogenesis of CFT073.
The CFT073 double mutant also exhibited consistently reduced bacterial titers in the
bladder on day 7 PI (Fig. 3.5C). Further investigation is required to determine the
significance of this observation.
To analyze further the kidney colonization defects of the UTI89 ΔhofQ and
CFT073 ΔyheF ΔhofQ deletion mutants in the CBA/J mouse model of ascending UTI, we
sectioned bladders and right side kidneys from animals inoculated with mutants, WT or
PBS alone for histological examination. The WT infected kidney sections showed various
signs of inflammation, while the inflammation in mutant infected kidney sections
correlated with the presence or absence of detectable CFUs in the corresponding left side
kidneys (Figs. 3.7 and 3.8). Since colonization of kidneys occurred at similar levels with
WT and mutant strains on day 1 PI, the mice with undetectable kidney CFUs on day 7 PI
51
were able to completely clear the mutant bacteria, with no signs of the prior infection.
Unfortunately, the histological examination of the infected kidneys did not provide
insights into the possible mechanism behind the defect of the mutants.
Overall, results from the animal infection experiments define a specific role for
T4P and/or T2SS effectors in the virulence of UPEC model strains UTI89 and CFT073.
52
IV. FIGURES
FIG. 3.1. Schematic representation of the T2SS and T4P gene clusters in UTI89 and
CFT073. The different T2SS and T4P gene clusters in the genomes of UTI89 and
CFT073 are shown. The detailed organization of the two T2SS gene clusters are shown in
the ovals. For the K12-T2SS, the E. coli K12 gene names are listed above the arrows and
the generic gsp nomenclature is given below. The OM secretin genes are indicated by the
hatched arrows. The open arrows are genes that do not show homology to any of the
known T2SS or T4P genes.
53
FIG. 3.2. Expression of the T2SS and T4P secretin genes. RNA was isolated from WT
strains of UTI89 (1), CFT073 (2), or MG1655 (3) grown in LB or tissue culture medium,
as indicated. RT-PCR was conducted using primer pairs (Table 1) specific for the OM
secretin genes hofQ, yheF or gspD. RT-PCR of the 16S rRNA gene was used as an
internal control to compare expression levels among the three strains.
54
FIG. 3.3. PCR verification of the different secretin knockout mutants. PCR was
performed on WT or knockout strains using primers flanking the indicated genes (Table
2.2). (A) UTI89 strains. Odd lanes, WT strain; even lanes, strains deleted for the
indicated genes. (B) CFT073 strains. Lanes 1 and 3, WT strain; lane 2, ∆yheF strain; lane
4, ∆hofQ strain; lanes 5 and 6, ∆yheF ∆hofQ strain.
55
FIG. 3.4. Comparison of growth curves of the OM secretin deletion mutants and
their WT counterparts. To confirm that the constructed deletion mutants did not exhibit
any growth defects, the mutants and the parental WT strains of UTI89 (A and B) and
CFT073 (C and D) were grown in LB (A and C) or RPMI + 10% FBS (B and D). Growth
was monitored by measuring the absorbance spectrophotometrically at 600 nm. Growth
of each mutant was similar to its WT counterpart.
56
FIG. 3.5. Mouse infection experiments. Eight week-old, female CBA/J mice were
infected by trans-urethral catheterization with WT or secretin knockout strains of UTI89
(A and B) or CFT073 (C and D). Bladders and kidneys were harvested on day 7 PI,
weighed, homogenized and CFU/g of organ weight determined. The UTI89 ΔhofQ
mutant was consistently defective in colonizing kidneys (p=0.03) (B), while no
differences were observed in CFU recovered from the bladder (A). In CFT073, the ∆yheF
∆hofQ double mutant was defective in colonizing the kidneys (p=0.013) (C), while
bladder CFU were slightly lowered as compared to WT (D). Median CFU values are
indicated by the line. Dotted line indicates the limit of detection (50 CFU/ g of kidney
and 500 CFU/g of bladder). The data were compared using Mann Whitney test for nonparametric data with one tailed P value. P<0.05 was considered significant.
57
FIG. 3.6. Complementation of the UTI89 and CFT073 mutants. Plasmid-borne copies
of yheF and hofQ were transformed into the mutants UTI89 ΔhofQ and CFT073 ΔyheF
ΔhofQ. 8-week old female CBA/J mice were infected with the mutants, corresponding
WT strains or the complemented strains by transurethral catheterization. On day 7 PI, the
bladders and kidneys were harvested from the infected animals, homogenized, and CFUs
enumerated. The complemented strains in UTI89 (A) as well as CFT073 (B) were able to
colonize and persist in the renal tissues just as well as the WT strains. Median CFU
values are indicated by the line. Dotted line indicates the limit of detection (50 CFU/ g of
kidney). The data were compared using Mann Whitney test for non-parametric data with
one tailed p value. p<0.05 was considered significant.
58
FIG. 3.7. Histological examination of the kidney sections from mice infected with
mutant and WT strains of UTI89. Kidneys obtained from mice inoculated with PBS
alone (A), UTI89 WT (B) or UTI89 ΔhofQ (C) were sectioned, H & E stained and
examined microscopically. Different regions in the kidney are indicated as follows; C;
calyx, CT; collecting tubule; RT; renal tubules, RP; renal papilla. The areas of leukocyte
invasion are indicated by arrows. The images are presented at 100X magnification and
the scale bar is 500 μm.
59
FIG. 3.8. Histological examination of the kidney sections from mice infected with
mutant and WT strains of CFT073. Kidneys obtained from mice inoculated with PBS
(A), CFT073 WT (B), or CFT073 ΔyheF ΔhofQ (C and D) were sectioned, H & E stained
and examined microscopically. Different regions in the kidney are indicated as follows;
C; calyx, CT; collecting tubule; RP; renal papilla. The areas of leukocyte invasion are
indicated by arrows. The images are presented at 100X magnification and the scale bar is
500 μm.
60
Chapter Four
Investigation of the Molecular Basis for the Virulence Defects
of UTI89 ΔhofQ and CFT073 ΔyheF ΔhofQ
I. INTRODUCTION
T2S and T4P systems are involved in the secretion and/or assembly of important
virulence factors in a variety of Gram-negative pathogenic bacteria. Both of the model
UPEC clinical strains- pyelonephritogenic isolate CFT073 and the cystitic isolate UTI89possess T2SS and T4P gene clusters homologous to those found in E. coli K12 (12, 104).
Moreover, UTI89 possesses a second T2SS gene cluster homologous to the ETEC-T2SS
(12). To understand the importance of these systems in the pathogenesis of UPEC, we
generated single and double deletion mutants in the OM secretin genes of CFT073 and
UTI89 and compared them with their parental WT strains using animal infection
experiments, as described in Chapter Three. These experiments showed that the kidney
CFUs recovered from mice infected with UTI89 ΔhofQ or CFT073 ΔyheF ΔhofQ mutant
strains were significantly lower than those of WT-infected mice (Fig. 3.7). This in turn
indicates that T4P and probably T2SS effectors play some role in the persistence of
UPEC in the upper urinary tract, especially the kidneys.
To gain further insight into the importance of T2SS and T4P in the virulence of
UPEC, we studied interactions of different secretin mutants and their parental WT strains
with human bladder and kidney epithelial cell lines, using a modified gentamicin
61
protection assay protocol (Materials and Methods) and examined the effects of the WT
and mutant bacteria on the cell lines. A panel of UPEC clinical isolates was obtained and
screened for the presence of the T2S and T4P systems to determine if the systems are
correlated with specific UTI disease. We also assessed secretion and assembly of T4P
indirectly by constructing UTI89 ΔfimD (type 1 pilus usher) deletion mutants in a ΔhofQ
or WT background. These deletion mutants were compared with ΔhofQ and WT UTI89
strains for their ability to form biofilms and to adhere to the monolayers of bladder and
kidney cell lines.
Overall, these experiments revealed a correlation between the UTI89 hofQ (T4P)
and gspD (ETEC-T2SS) genes and ability to efflux out of bladder epithelial cells, but did
not identify specific mechanisms to account for the virulence attenuation observed in
vivo.
62
II. RESULTS
Interactions of the UPEC secretin mutants with bladder and kidney
epithelial cells. The function of the T2SS and T4P in UTI89 and CFT073 was first
assessed using a modified gentamicin protection assay (see Materials and Methods, (61)).
This assay tests the ability of UPEC to adhere to and invade human urothelial cells, form
intracellular replicating foci, and efflux out of the infected cells. The assay was carried
out using the 5637 human bladder and A498 human kidney cell lines. None of the
secretin knockout mutants in CFT073 or UTI89 exhibited defects in adhering to (Fig. 4.1)
or invading (Fig. 4.2) the human bladder or kidney cells compared to the WT strains.
Bladder epithelial cells are exfoliated and washed off by urine as an innate
immune response to infection by UPEC (69). To escape the exfoliating cells, UPEC
efflux out of the infected cells and spread to other parts of the urinary tract. None of the
CFT073 secretin mutants exhibited defects in effluxing out of the bladder or kidney cells
(data not shown). In addition, the UTI89 secretin knockout mutants effluxed out of
kidney cells similar to the WT strain. However, the UTI89 ∆hofQ (T4P) and ∆gspD
(ETEC-T2SS) strains consistently exhibited approximately 2-fold reductions (p<0.01) in
their ability to efflux out of human bladder cells as compared to the WT UTI89 strain
(Fig. 4.3). Thus, both the ETEC-T2SS and T4P are required for efficient escape from
bladder cells, at least for the UTI89 cystitis isolate.
An intracellular growth experiment for an extended period of 48 hours (as
explained in Materials and Methods) was carried out. No significant differences were
observed among the intracellular growth patterns of any of the OM secretin deletion
mutants and the WT strains of UTI89 for the 5637-bladder cell line (Fig. 4.4). Moreover,
63
no difference was observed between the intracellular growth index (IGI) values for any of
the deletion mutants and parental WT strain of UTI89 during the 6 h efflux period (Data
not shown). These results established that the efflux defective nature of ΔgspD and
ΔhofQ mutants in UTI89 is not a manifestation of defective growth inside the bladder
epithelial cells.
Because the effluxed bacteria remain in the tissue culture medium for the 5637
bladder epithelial cell line for a period of 6 h during the efflux assay, we also compared
the growth of the ΔgspD and ΔhofQ mutant strains with WT UTI89 in RPMI + 10% FBS
(Fig. 3.4B); the mutants did not exhibit any growth defects, ruling out that the decreased
efflux index values were due to lack of growth of the mutants in the tissue culture
medium.
Effects of UPEC secretin mutants on IL-6 production and cytotoxicity. UPEC
has been shown to suppress the induction of IL-6, a pro-inflammatory and
immunomodulatory cytokine, produced by the urothelium in response to bacterial LPS
and IL-1 (45). One hypothesis explaining the mechanism behind the cytokinesuppression is that UPEC secretes effectors that downregulate host cell NF-κB signaling
to repress transcription of pro-inflammatory and apoptotic genes (45). To analyze
whether T2SS and/or T4P are involved in the secretion of this unknown effector, we
compared the different UTI89 and CFT073 knockout mutants with their parental WT
strains for ability to suppress IL-6 production in response to LPS by the 5637 human
bladder cell line. The knockout mutants were able to suppress IL-6 formation just as well
as the WT strains (data not shown). We also tested whether the T2SS and/or T4P might
be involved in the secretion of a toxic effector molecule. The UTI89 and CFT073
64
secretin knockout mutants were compared with their parental WT strains for differences
in cytotoxicity toward the 5637 human bladder cell line. None of the mutants exhibited
decreased cytotoxicity, as measured by LDH release during a 3-hour infection assay (Fig.
4.5). In addition, no differences were detected between WT UTI89 and the secretin
knockout mutants in the amount of LDH released during the efflux assay carried out as
described above (Fig. 4.6). This indicates that a defect in secretion of a cytotoxic effector
molecule is not responsible for the decreased efflux of the UTI89 ∆hofQ and ∆gspD
strains from bladder epithelial cells.
Prevalence of T2SS and T4P in a panel of pathogenic and non-pathogenic E.
coli. The cystitic isolate, UTI89, possesses a second T2SS gene cluster homologous to
one found in ETEC. In ETEC, the T2SS is known to be involved in the secretion of an
important ETEC virulence factor, LT (94). Furthermore, the UTI89 ΔgspD deletion strain
that was constructed in order to ablate the ETEC-T2SS was found to be defective in
fluxing out of the 5637, human bladder carcinoma cell line (Fig. 4.3). Based on these
findings we hypothesized that a correlation may exist between the presence of ETECT2SS in a UPEC isolate and the cystitic nature of that isolate.
We analyzed a panel of 58 different pathogenic and nonpathogenic E. coli strains
(Table 2.1) by colony PCR using primers specific for ETEC-T2SS and T4P genes. The
results from these experiments indicated that the ETEC-T2SS is widely distributed
among pathogenic E. coli without any relation to the origin or tropism of a clinical
isolate. Thus, the ETEC-T2SS does not appear to be specifically associated with cystitis.
Furthermore, all the 58 isolates were observed to possess the three T4P genes, hofQ,
yggR and hofB, that we tested for.
65
Next we checked whether an efflux-defect from bladder epithelial cells is directly
related to the absence of the ETEC-T2SS. We analyzed strains from the panel that were
negative by PCR for the ETEC-T2SS for their ability to flux out of monolayers of the
5637 human bladder epithelial cell line. From the results of this experiment, UPEC
strains lacking the ETEC-T2SS do not appear to be generally efflux-defective with
respect to the human bladder cell line (Fig. 4.7).
Analyzing the role of T4P in UTIs. As described earlier, T4P are long, surfaceassociated filaments that are involved in different virulence functions such as adhesion,
twitching motility, biofilm formation, and horizontal genetic transfer. Since, the ΔhofQ
mutant lacking the T4P OM secretin in UTI89, was defective in fluxing out of the human
bladder epithelial cell line as well as in renal persistence in the CBA/J mouse model of
ascending UTI, we decided to analyze this mutant further for any changes in the
functions associated with T4P such as adhesion and biofilm formation.
The two well-characterized adhesins used by UPEC to bind to bladder and kidney
epithelial cells are the FimH and PapG tips of type 1 and P pili, respectively. Moreover,
the adhesion frequency of a fimH deletion mutant for 5637 bladder epithelial cells drops
to negligible levels as compared to that of WT (61). Thus, FimH is the major adhesin that
allows UPEC to adhere to and later invade the bladder epithelial cells. In many Gramnegative pathogens, T4P have been shown to play an important role as the major
adhesive factors. However, ΔhofQ mutant showed no defect in adhesion to human
bladder or kidney epithelial cells. To check whether the type 1 pili were masking the
adhesive properties of the T4P, we generated type 1 pilus negative mutants in UTI89 by
targeting the type 1 usher fimD for deletion in the WT as well as ΔhofQ backgrounds.
66
These two mutant strains, ΔfimD and ΔfimD ΔhofQ, were compared with the WT UTI89
for adherence to the monolayers of human bladder (5637) as well as kidney (A498) cell
lines. As expected, the adhesion frequency of the ΔfimD mutant was lowered >20 times
for both the cell lines. Quite interestingly, the double deletion mutant, ΔfimD/ΔhofQ was
~50% less adherent than ΔfimD with respect to both bladder and kidney epithelial cells
(Fig 4.8 A and B). These results were consistent and statistically significant
Biofilms constitute a viscous phase composed of bacterial cells in consortium
with other microorganisms and bacterial exopolysaccharide matrices, within which
molecules and ions can occur at concentrations different from those found in the bulk
phase (14). UPEC form replication foci with biofilm-like properties within the bladder
epithelial cells (3), although nothing is known about their ability to form extracellular
biofilms. To analyze biofilm formation by both UTI89 and CFT073 and to understand the
role of T4P in this process, we compared different secretin deletion mutants with their
WT counterparts for the ability to form biofilms. The mutant and WT bacteria were
grown in LB, tissue culture medium (RPMI + 10% FBS), or M9 minimal medium in 96well polystyrene plates at room temperature or at 370C in the presence of 5% CO2.
Formation of biofilms was determined by staining with crystal violet. Both UTI89 and
CFT073 did not form biofilms in either tissue culture medium or LB. In contrast, strong
biofilm formation was detected in M9 minimal medium, but no differences were detected
between the biofilms formed by different deletion mutants and their parental WT strains
(Fig. 4.9). Thus, T4P do not appear to contribute to biofilm formation by UPEC, at least
under the in vitro conditions tested.
67
III. DISCUSSION
The cystitic strain UTI89 and the pyelonephritogenic strain CFT073, the two
model UPEC isolates, possess chromosomal copies of T2SS and T4P gene clusters. To
study the importance of these secretion pathways in the virulence of UPEC we generated
OM secretin deletion mutants and as described in Chapter Three established a
relationship between the presence of T4P and/or T2SS and the ability of UPEC to
colonize and persist in the kidneys of CBA/J mouse model of ascending UTI. In this
chapter, I have described experiments carried out to understand the mechanism governing
this relationship.
Using a modified gentamicin protection assay, we compared different secretin
deletion mutants with their WT counterparts for different aspects of uropathogenesis such
as adhesion to and entry into cells of human bladder and kidney epithelia, intracellular
replication and bacterial efflux out of the infected cells. Two deletion mutants in UTI89,
ΔhofQ and ΔgspD were found to be defective in fluxing out of the 5637 human bladder
epithelial cell line. Moreover, these deletion mutants were not defective in intracellular
growth, ability to grow in the tissue culture medium, cytotoxicity or IL-6 suppression.
This suggests that both T4P and the ETEC-T2SS contribute to the ability of UTI89 to
efflux out of the bladder epithelial cells.
The results from the semi-quantitative RT-PCR indicating that the ETEC-T2SS
was strongly expressed in vitro (Fig. 3.2), along with the results from the efflux assay and
the fact that the ETEC-T2SS is absent from the genome of the CFT073 pyelonephritis
isolate, raised the possibility that the ETEC-T2SS might allow secretion of a toxin or
other virulence factor by UTI89. This also suggested that the ETEC-T2SS might be
68
specifically associated with the ability to cause cystitis. However, results from the
analysis a panel of pathogenic and non-pathogenic E. coli isolates by DNA-PCR using
ETEC-T2SS specific primer pairs indicated that the ETEC-T2SS gene cluster is
widespread among E. coli clinical isolates, without any apparent correlation with the
origin of the isolate. Thus, CFT073 appears to be an outlier among E. coli strains in
lacking the ETEC-T2SS.
To examine the role of T4P in UTI89, we generated ΔhofQ deletion mutants in
WT as well as ΔfimD (type 1 pilus usher) backgrounds. In the absence of type 1 pili
(ΔfimD), the T4P seem to play a role in the adhesion of UTI89 to human bladder as well
as kidney cell lines. The adhesion frequency values for the ΔfimD ΔhofQ double deletion
mutants in 5637 bladder and A498 kidney epithelial cells, were consistently (~50%)
lower than that for UTI89 ΔfimD strains (Fig. 4.8). Although this difference is
statistically significant, the actual adhesion frequency values are so low (0.5-0.6%) that
the role played by T4P in UPEC adherence to bladder and kidney epithelial cells can at
best be described as minor. Ability to form biofilms is a property often associated with
T4P. While CFT073 and UTI89 were able to form biofilms in M9 minimal medium, none
of the deletion mutants exhibited reduced biofilm formation as compared to the parental
WT strains (Fig. 4.9). This indicates that in UPEC the ability to form biofilms is not
dependent on the presence of functional T4P. Note that these experiments also found no
biofilm defects in strains containing deletions in T2SS or type 1 pili.
69
IV. FIGURES
FIG. 4.1. Adhesion of secretin deletion mutants and WT strains to bladder and
kidney epithelial cell lines. At the end of the two hour period of adhesion in the
modified gentamicin protection assay, the infected cell lines were washed vigorously and
the adherent bacterial counts enumerated. Different secretin deletion mutants in UTI89
(A and B) or CFT073 (C and D) were able to adhere to 5637 human bladder and A498
human kidney epithelial cell lines as well as their WT counterparts. Percent adhesion
frequency values are indicated on the ordinate. All data are representative of three or
more assays, each performed in triplicate.
70
FIG. 4.2. Invasion of secretin deletion mutants and WT strains into bladder and
kidney epithelial cell lines. In the modified gentamicin protection assay, following a
period of adhesion, the infected cell lines were treated with 100 μg/ml gentamicin for two
hours and the intracellular CFUs at the end of this period determined. Different secretin
deletion mutants in UTI89 (A and B) or CFT073 (C and D) were able to invade the 5637
human bladder and A498 human kidney epithelial cell lines to WT-like levels. Percent
invasion frequency values are indicated on the ordinate. All data are representative of
three or more assays, each performed in triplicate.
71
FIG. 4.3. Efflux of WT UTI89 and the indicated UTI89 secretin mutants from 5637
human bladder epithelial cells. The UTI89 ΔgspD and ΔhofQ mutants consistently
exhibited significantly reduced efflux index values compared to bladder cells infected
with WT UTI89. Efflux index values for the UTI89 ∆yheF mutant were not always
reduced compared to WT. Each deletion mutant was compared with the WT strain at
least 3 times, each in triplicate. Values are shown as a percentage of WT; error bars
indicate SD. The efflux index values were compared by Dunnett’s multiple comparison’s
test. *, p<0.05; **, p<0.01.
72
FIG. 4.4. Intracellular replication of secretin deletion mutants and parental WT
strains of UTI89. The OM secretin deletion mutant and WT UTI89 strains were allowed
to grow inside 5637 bladder cells over a period of 48 h. 10μg/ml gentamicin was
maintained in the surrounding medium. At indicated times, the infected cells were
washed, treated with TX100 and intracellular CFUs determined. The OM secretion
deletion mutants in UTI89 were able to replicate inside 5637 human bladder cells in a
WT-like manner.
73
FIG. 4.5. Instant cytotoxicity of UTI89 (A) or CFT073 (B) for human bladder
epithelial cells. WT strains or the different secretin knockout mutants were tested for
their ability to cause cytotoxicity in 5637 human bladder cells over a period of 3 h, as
measured by LDH release (% cytoxicity). Bacteria were added at an MOI of 5 or 10.
74
FIG. 4.6. Cytotoxicity caused by UTI89 during the 6 h efflux period. The
supernatants were collected at indicated time points during the efflux assay carried out by
infecting 5637 bladder epithelial cells with the T2S and T4P OM secretin deletion
mutants and WT strains of UTI89. The cytotoxicity was assessed by estimating LDH
released in the cell-free supernatants.
75
FIG. 4.7. Efflux index values for the Representative ETEC-T2SS negative strains. In
the 6 h efflux assay module, representative strains from each clinical group of the
pathogenic E. coli panel that were negative for ETEC-T2SS specific PCR did not exhibit
defects in fluxing out of the monolayers of 5637 human bladder cell line. Error bars
indicate SD.
76
FIG. 4.8. Role of T4P in the adhesion to human bladder and kidney epithelial cells.
Percent adhesion frequency for 5637 bladder (A) and A498 kidney (B) cells was
determined for hofQ deletion mutants in WT as well as fimD deletion backgrounds in
UTI89. All date are representative of two or more assays, each performed in triplicate.
The results were compared by unpaired t test with one-tail P value. p<0.05 was
considered significant.
77
FIG. 4.9. Biofilm formation in M9 medium. Different OM secretin deletion mutants
and parental WT strains of UTI89 (A) and CFT073 (B) were tested for their ability to
form biofilms in M9 minimal medium in quadruplicate. Ps indicates the biofilm formed
by P. aeruginosa. OD575 ≥ 0.8 is considered positive for biofilm formation. Each
histogram represents the average of four OD575 values +/- SD. Based on these results, the
biofilm formation in UPEC strains UTI89 and CFT073 does not appear to be dependent
on presence of T4P, T2SS or type 1 pili.
78
Chapter Five
Discussion and Future Directions
I. CONCLUSIONS
Many Gram-negative bacterial pathogens use type II secretion and the
evolutionarily related T4P system to secrete and/or assemble virulence factors. Although
functions for these secretion pathways are established for intestinal pathogenic E. coli,
their roles in the virulence of ExPEC had not been investigated. In this report, we
addressed this question by constructing deletion mutations to disable T2SS and T4P gene
clusters present in the genomes of the model UPEC strains UTI89 and CFT073. We
compared these mutants to their WT counterparts using different in vitro and in vivo
assays. Our results demonstrate that T4P and potentially also the K12-T2SS are important
to the pathogenesis of UPEC in the urinary tract, particularly in the ability to successfully
colonize and persist within the kidneys.
Both the cystitis isolate UTI89 and the pyelonephritis isolate CFT073 possess
T2SS and T4P gene clusters sharing very close homology (generally 91-99% amino acid
identity) with genes present in E. coli K12. In K12 strains, the T2SS does not appear to
be expressed under standard growth conditions, but this system was shown to be derepressed in mutants lacking the nucleoid structuring protein H-NS and to function in the
secretion of chitinase (28). Similarly, the E. coli K12 T4P genes are transcribed at low
levels and T4P are not detected under standard laboratory conditions (87). However, the
major pilin subunit PpdD (prepilin peptidase dependent protein D) was assembled into
79
T4P when ectopically expressed in P. aeruginosa (87). We found that the K12homologous T2SS and T4P genes in UTI89 and CFT073 were expressed at low levels
under in vitro culture conditions, similar to expression in a K12 strain. Since our infection
studies indicate that these systems are functional in vivo, they are likely to be upregulated
under yet-to-be-determined conditions. Notably, our results imply that the T4P and T2SS
genes in E. coli K12 are also functional under the proper conditions.
Interestingly, in addition to the K12-homologous T2SS and T4P genes, the UTI89
cystitis isolate contains a second T2SS cluster homologous to genes present in ETEC. In
ETEC this T2SS enables the secretion of LT, which is critical for pathogenesis in the
intestinal tract (94). In contrast to the K12-homologous systems, we found that the
ETEC-T2SS was strongly expressed in vitro. This raised the possibility that the ETECT2SS might allow secretion of a toxin or other virulence factor by UTI89. Furthermore,
since the ETEC-T2SS is absent from the genome of the CFT073 pyelonephritis isolate,
and because the UTI89 mutant lacking a functional ETEC-T2SS (ΔgspD) was
specifically defective in fluxing out of the bladder epithelial cells, this suggested that the
ETEC-T2SS might be specifically associated with the ability to cause cystitis. To check
for a correlation between the cystitic nature of a UPEC strain and the presence of the
ETEC-T2SS, we analyzed a panel of pathogenic and non-pathogenic E. coli isolates by
PCR using ETEC-T2SS specific primer pairs. Results from this analysis indicated that
the ETEC-T2SS gene cluster is widespread among E. coli clinical isolates, without any
apparent correlation with the origin of the isolate. This observation was further
corroborated by examination of the available genome sequences of E. coli strains by
80
BLAST analysis. Thus, CFT073 appears to be an outlier among E. coli strains in lacking
the ETEC-T2SS.
In the CBA/J mouse model of ascending UTI, we found that the ΔhofQ (T4P)
mutant of UTI89 was significantly defective in colonization of kidneys on day 7 PI as
compared to WT UTI89, although there was no apparent defect in the ability of this strain
to colonize and persist within the bladder. In addition, we did not detect any defects in the
abilities of the UTI89 ∆yheF (K12-T2SS) or ∆gspD (ETEC-T2SS) mutants to colonize
the bladder or kidneys at any time points. Thus, our experiments revealed a specific role
for T4P in the ability of UTI89 to persistently colonize the kidney. In the
pyelonephritogenic isolate CFT073, mutants containing single deletions of hofQ or yheF
were able to infect and colonize bladder and kidney tissues just as well as the WT strain.
However, similar to the UTI89 ∆hofQ strain, a CFT073 double knockout mutant lacking
both yheF and hofQ showed significantly reduced colonization of the kidney on day 7 PI.
The CFT073 double mutant also exhibited consistently reduced bacterial titers in the
bladder on day 7 PI. These results demonstrate that T4P and/or the T2SS play important
roles in the ability of CFT073 to persistently colonize the urinary tract. Further
investigation is required to determine the specific contributions of T4P and the T2SS to
the pathogenesis of CFT073. OM secretins are capable of functioning in more than one
secretion pathway within a bacterium. For example, in V. cholerae, the EpsD secretin is
able to secrete proteins as part of a T2SS while at the same time helping in the extrusion
of filamentous CTX phage (84). In the case of CFT073, the two OM secretins, YheF and
HofQ, might be able to replace one another functionally and hence the colonization
81
defect was observed only when the mice were infected with the double knockout mutant
lacking both secretins.
In vitro experiments using model cell lines do not reflect the myriad aspects of
host-pathogen interaction and our experiments to analyze interactions of the UTI89 and
CFT037 mutants with human bladder and kidney epithelial cells did not point to a
mechanism underlying the defects observed in the mouse infection experiments. UPEC
adhere to cells of the host urothelium with the help of the type 1 and P pili, and in case of
the bladder epithelium their entry into the superficial umbrella cells is mediated by the
type 1 pilus tip adhesin, FimH (61). Hence it was expected that none of the knockout
mutants, generated by specifically targeting the type II secretion and T4P pathways,
would be defective in adhesion or invasion; although in the absence of type 1 pili
(ΔfimD), T4P played a minor role in the adhesion of UTI89 to cells of the bladder and
kidney epithelium. In addition, we did not detect defects in intracellular replication,
cytotoxicity, or suppression of IL-6 expression by bladder epithelial cells. Using a
modified gentamicin protection assay that mimics several aspects of UPEC pathogenesis,
we did observe that knockout mutants in UTI89 lacking hofQ (T4P) and gspD (ETECT2SS) were defective in fluxing out of the bladder epithelial cells line. However, this
defect in fluxing out of the epithelial cells did not correlate with the behavior of the
mutants in the mouse UTI model, where only the ΔhofQ mutant exhibited decreased
colonization. Observation of green fluorescent protein-expressing UTI89 in the cell
culture efflux assay confirmed defects in the ΔhofQ and ΔgspD mutants, but indicated a
delay in efflux rather than an absolute defect (data not shown). This could explain the
lack of correlation to the in vivo results. Moreover, the CFT073 secretin mutants did not
82
exhibit any defects in fluxing out of bladder epithelial cells. Finally, the progression of
UTI in the kidney is not documented in as many details as in the bladder. From our
experiments, it is apparent that all knockout mutants in UTI89 as well as in CFT073 were
able to form IBCs in the A498 human kidney cell line.
Based on my dissertation work, we propose that T4P aid in the colonization and
persistence of UPEC in the urinary tract, particularly the upper urinary tract. The ascent
of UPEC along the ureters is not well understood and different bacterial factors that
might aid in this process are still undefined. In a number of Gram-negative pathogens,
association with the host epithelium followed by motility and dispersal over the epithelial
surface are mediated by T4P (63). The T4P in UPEC might play a similar role, although
colonization of the urinary tract at early time points PI were similar in both the mutant
and WT strains. Alternatively, T4P may specifically aid in the colonization of kidney
tissues by UPEC, possibly by mediating bacterial-bacterial and bacterial-host interactions
and/or the formation of biofilms. Another possibility is that in the absence of functional
T4P and/or the T2SS, the mutant bacteria were defective due to the lack of an unknown
secreted effector, leading to clearance of the bacteria by day 7 PI.
83
II. FUTURE DIRECTIONS
My dissertation research established a correlation between T4P and/or effectors of
the T2SS with the ability of UPEC to colonize and persist in renal tissue. However, we
do not currently understand the detailed mechanism behind this relationship. Future
research should be directed toward elucidating the molecular basis of the virulence
attenuation of the UPEC mutants. This will further our understanding of E. coli
pathogenesis within the urinary tract. Potential future experiments are described below.
Study of the regulation of T2SS and T4P proteins in UPEC. As described in
Chapter Three, the results from the semiquantitative RT-PCR experiments revealed that
the K12-T2SS as well as T4P genes were transcribed at low levels, while the ETECT2SS was transcribed at relatively high levels under the in vitro conditions tested. To
study the expression of the UPEC T2SS and T4P in vivo, and to understand the temporal
aspects of their regulation, transcription fusion constructs to reporters such as βgalactosidase or luciferase could be generated. A new technique to generate specific
reporter gene fusions in bacterial chromosomes is described by Gerlach, et al (32). This
technique is very similar to the method used for generation of the secretin deletion
mutants. It makes use of bacteriophage λ RED recombinase to insert a PCR-generated
gene cassette containing a promoterless reporter gene together with a FRT-flanked
antibiotic resistance cassette behind the ribosome binding site and the start codon of the
target gene. These T2SS and T4P reporter fusions can be used in in vitro cell culture
assays and in vivo animal model infection experiments to monitor gene expression. Once
the correct condition(s) for the induction of T2S and T4P systems are identified,
additional experiments such as analysis of secretion profiles of mutants and WT strains or
84
the visualization of T4P on the surface of UPEC by EM can be carried out. These
experiments may also aid in understanding whether the K12 T2SS or T4P is more
important in the renal colonization and persistence of CFT073.
In vivo imaging of host-UPEC interactions. The ability to monitor the progress
of infections in the mouse model will greatly aid understanding of the function of the T2S
and T4P systems in UPEC pathogenesis. A first approach to comparing the WT and
mutant strains during the course of an infection is to conduct detailed histological
analysis during different time points post infection. A powerful new approach is the field
of in vivo imaging.
One method for this employs pathogens that are genetically
engineered to express the lux operon from bacteria such as Photorhabdus luminescence
(46). Although this technique has not been employed and standardized for research
related to UPEC mediated UTI, there is ample precedent for the use of bioluminescent in
vivo imaging in the study of infections caused by pathogenic bacteria such as Salmonella
typhimrium (13), Streptococcus pneumoniae (30), viruses such as herpes simplex virus
(HSV)-1 (58), and even protozoan parasites such as Toxoplasma gondii (43). The lux
operon encodes genes to synthesize the enzyme luciferase and its substrate luciferin.
Thus the pathogenic bacteria (UPEC) will constitutively emit light without need for an
exogenous substrate. UPEC engineered to express lux (WT as well as OM secretin
deletion mutants) can be used to infect the CBA/J mice by transurethral catheterization
and the disease progression monitored by placing the infected mice under a chargecoupled device (CCD) camera that can detect the low levels of luciferase emitted by the
UPEC in vivo. The in vivo imaging procedure using the bioluminescent UPEC will allow
the progression of UTI in CBA/J mice to be monitored without killing the animals and to
85
gain spatial and temporal insights into the colonization of the upper urinary tract,
especially the renal tissue, by the ΔhofQ deletion strain in UTI89, the ΔyheF ΔhofQ
double deletion mutant in CFT073 and their WT counterparts.
Role of T2S and T4P systems in bladder infections caused by UPEC. As
described in Chapter One, the UPEC developmental cascade in the bladder consists of
bacterial adherence to the bladder epithelium followed by invasion, establishment and
maturation of intracellular replicating foci and emergence of filamentous bacteria from
the infected cells. The data from the in vitro bladder epithelial cell infection experiments
showed that T4P and ETEC-T2SS are important for the efflux of UTI89 from bladder
cells. However, no defects in bladder colonization and persistence were observed in
animal infections with UTI89 mutants containing deletions in ETEC-T2SS or T4P.
Further analysis of these mutants can be carried out by histological analysis of bladder
samples obtained at different time points PI. The detailed study of the passage of UPEC
through the maturation cascade in the bladder has been observed by the time lapse videomicroscopic analysis of murine bladder mounted ex vivo on an anodized aluminum
incubation chamber (50). This method can be used to analyze T2SS and T4P secretin
deletion mutants for defects in steps of the developmental cascade.
Importance of T2S and T4P systems in intestinal colonization. UPEC (and
other ExPEC) do not cause gastro-intestinal disease in humans, although they are able to
establish prolonged intestinal colonization, at times even more effectively than the
intestinal pathogenic E. coli (49). UPEC typically enter the urinary tract as a consequence
of fecal contamination of the urethral opening and cause ascending UTI. Thus, UPEC
must be able to successfully pass through the gastro-intestinal tract. As described earlier,
86
T2SS effectors such as LT in the case of ETEC or the bundle forming T4P of EPEC are
known to play important roles in diseases caused by intestinal pathogenic E. coli (33, 34,
94). Moreover, UTI89 possesses a T2SS gene cluster that is homologous to one found in
ETEC. Based on this information, it will be quite interesting to check whether the T2S
and T4P systems in the model UPEC strains, UTI89 and CFT073, play a role in their
ability to colonize and pass through the intestine. To address this, OM secretin mutants in
which the antibiotic resistant cassette has not been removed can be used to infect mice
orally along with their parental WT strains (marked by a different antibiotic resistance).
The comparison of CFUs recovered from the feces of animals infected with different
mutant and WT strains will indicate whether the lack of T2S and/or T4P systems has any
effect on the ability of UPEC to colonize the intestine.
Concluding Remarks. ExPEC is an inclusive designation for a distinctive group
of pathogenic E. coli based on their common virulence factors, serotypes and
phylogenetic groups, and includes mainly the pathogenic strains causing UTI,
sepsis/neonatal meningitis in humans and extra-intestinal (e.g.; respiratory) infection in
poultry. ExPEC employ a broad array of virulence factors, including type 1 and P pili,
and secreted toxins such as CNF1, hemolysin and Sat. Using the model UPEC isolates
CFT073 and UTI89 we sought to understand the importance of the type II secretion and
type IV pilus systems in the virulence of UPEC. The deletion mutants UTI89 ΔhofQ and
CFT073 ΔyheF ΔhofQ showed reduced ability to colonize and persist in the upper urinary
tract, especially kidneys. This defect was rescued by plasmid-borne copies of the specific
87
genes. These results fulfill the molecular Koch’s postulates, in turn emphasizing that T4P
are important virulence factors of UPEC and T4P and/or T2SS effectors aid in the
colonization and persistence of UPEC in the upper urinary tract. Our results expand the
impressive list of weaponry used by ExPEC to include T4P and possibly also the T2SS
and open new avenues for understanding the molecular mechanisms of E. coli
pathogenesis in the upper urinary tract.
88
Abbreviations
APEC
AUM
Cdi
CFU
CNF1
ETEC
ExPEC
FBS
GSP
GTP
H&E
HlyA
IBC
IM
KO
LB
LDH
LT
MEM
NMEC
OD
OM
Pap
PI
PMN
PS
QIR
Sat
T2SS
T4P
TLR
TX 100
UP
UPEC
UTI
WT
Avian Pathogenic Escherichia coli
Asymmetric unit membrane
Contact-dependent inhibitor
Colony forming unit
Cytotoxic necrotizing factor 1
Enterotoxigenic Escherichia coli
Extra-intestinal Pathogenic Escherichia coli
Fetal bovine serum
General secretory pathway
Guanosine tri phosphate
Hematoxylin and eosin
Hemolysin A
Intracellular bacterial community
Inner membrane
Knock out
Luria Broth
Lactate dehydrogenase
Heat-labile enterotoxin
Modified Eagle’s medium
Neonatal Meningitis-causing Escherichia coli
Optical density
Outer membrane
Pyelonephritis-associated pili
Post-inoculation
Polumorphonuclear leukocyte
Protamine sulphate
Quiescent intracellular reservoir
Secreted autotransporter protein
Type II secretion system
Type IV pilus
Toll-like receptor
Triton X 100
uroplakin
Uropathogenic Escherichia coli
Urinary Tract Infections
Wild type
89
References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
Alm, R. A., and J. S. Mattick. 1997. Genes involved in the biogenesis and
function of type-4 fimbriae in Pseudomonas aeruginosa. Gene 192:89-98.
Anderson, G. G., K. W. Dodson, T. M. Hooton, and S. J. Hultgren. 2004.
Intracellular bacterial communities of uropathogenic Escherichia coli in urinary
tract pathogenesis. Trends Microbiol 12:424-30.
Anderson, G. G., J. J. Palermo, J. D. Schilling, R. Roth, J. Heuser, and S. J.
Hultgren. 2003. Intracellular bacterial biofilm-like pods in urinary tract
infections. Science 301:105-7.
Aoki, S. K., R. Pamma, A. D. Hernday, J. E. Bickham, B. A. Braaten, and D.
A. Low. 2005. Contact-dependent inhibition of growth in Escherichia coli.
Science 309:1245-8.
Blattner, F. R., G. Plunkett, 3rd, C. A. Bloch, N. T. Perna, V. Burland, M.
Riley, J. Collado-Vides, J. D. Glasner, C. K. Rode, G. F. Mayhew, J. Gregor,
N. W. Davis, H. A. Kirkpatrick, M. A. Goeden, D. J. Rose, B. Mau, and Y.
Shao. 1997. The complete genome sequence of Escherichia coli K-12. Science
277:1453-74.
Bleves, S., M. Gerard-Vincent, A. Lazdunski, and A. Filloux. 1999. Structurefunction analysis of XcpP, a component involved in general secretory pathwaydependent protein secretion in Pseudomonas aeruginosa. J Bacteriol 181:4012-9.
Bock, K., M. E. Breimer, A. Brignole, G. C. Hansson, K. A. Karlsson, G.
Larson, H. Leffler, B. E. Samuelsson, N. Stromberg, C. S. Eden, and et al.
1985. Specificity of binding of a strain of uropathogenic Escherichia coli to Gal
alpha 1-4Gal-containing glycosphingolipids. J Biol Chem 260:8545-51.
Bose, N., S. M. Payne, and R. K. Taylor. 2002. Type 4 pilus biogenesis and type
II-mediated protein secretion by Vibrio cholerae occur independently of the
TonB-facilitated proton motive force. J Bacteriol 184:2305-9.
Bower, J. M., D. S. Eto, and M. A. Mulvey. 2005. Covert operations of
uropathogenic Escherichia coli within the urinary tract. Traffic 6:18-31.
Bruijn, F. J. d., Lupski, James R, Weinstock George M, . 1998. Bacterial
Genomes Physical Structure and Analysis, vol. Springer.
Burrows, L. L. 2005. Weapons of mass retraction. Mol Microbiol 57:878-88.
Chen, S. L., C. S. Hung, J. Xu, C. S. Reigstad, V. Magrini, A. Sabo, D.
Blasiar, T. Bieri, R. R. Meyer, P. Ozersky, J. R. Armstrong, R. S. Fulton, J.
P. Latreille, J. Spieth, T. M. Hooton, E. R. Mardis, S. J. Hultgren, and J. I.
Gordon. 2006. Identification of genes subject to positive selection in
uropathogenic strains of Escherichia coli: a comparative genomics approach. Proc
Natl Acad Sci U S A 103:5977-82.
Contag, C. H., P. R. Contag, J. I. Mullins, S. D. Spilman, D. K. Stevenson,
and D. A. Benaron. 1995. Photonic detection of bacterial pathogens in living
hosts. Mol Microbiol 18:593-603.
90
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
Costerton, J. W., K. J. Cheng, G. G. Geesey, T. I. Ladd, J. C. Nickel, M.
Dasgupta, and T. J. Marrie. 1987. Bacterial biofilms in nature and disease.
Annu Rev Microbiol 41:435-64.
Craig, L., R. K. Taylor, M. E. Pique, B. D. Adair, A. S. Arvai, M. Singh, S. J.
Lloyd, D. S. Shin, E. D. Getzoff, M. Yeager, K. T. Forest, and J. A. Tainer.
2003. Type IV pilin structure and assembly: X-ray and EM analyses of Vibrio
cholerae toxin-coregulated pilus and Pseudomonas aeruginosa PAK pilin. Mol
Cell 11:1139-50.
Datsenko, K. A., and B. L. Wanner. 2000. One-step inactivation of
chromosomal genes in Escherichia coli K-12 using PCR products. Proc Natl Acad
Sci U S A 97:6640-5.
Davis, J. M., H. M. Carvalho, S. B. Rasmussen, and A. D. O'Brien. 2006.
Cytotoxic necrotizing factor type 1 delivered by outer membrane vesicles of
uropathogenic Escherichia coli attenuates polymorphonuclear leukocyte
antimicrobial activity and chemotaxis. Infect Immun 74:4401-8.
Davis, J. M., S. B. Rasmussen, and A. D. O'Brien. 2005. Cytotoxic necrotizing
factor type 1 production by uropathogenic Escherichia coli modulates
polymorphonuclear leukocyte function. Infect Immun 73:5301-10.
Eto, D. S., T. A. Jones, J. L. Sundsbak, and M. A. Mulvey. 2007. IntegrinMediated Host Cell Invasion by Type 1-Piliated Uropathogenic Escherichia coli.
PLoS Pathog 3:e100.
Eto, D. S., J. L. Sundsbak, and M. A. Mulvey. 2006. Actin-gated intracellular
growth and resurgence of uropathogenic Escherichia coli. Cell Microbiol 8:70417.
Ewers, C., G. Li, H. Wilking, S. Kiessling, K. Alt, E. M. Antao, C. Laturnus,
I. Diehl, S. Glodde, T. Homeier, U. Bohnke, H. Steinruck, H. C. Philipp, and
L. H. Wieler. 2007. Avian pathogenic, uropathogenic, and newborn meningitiscausing Escherichia coli: how closely related are they? Int J Med Microbiol
297:163-76.
Falkow, S. 2004. Molecular Koch's postulates applied to bacterial pathogenicity-a personal recollection 15 years later. Nat Rev Microbiol 2:67-72.
Falkow, S. 1988. Molecular Koch's postulates applied to microbial pathogenicity.
Rev Infect Dis 10 Suppl 2:S274-6.
Forest, K. T., and J. A. Tainer. 1997. Type-4 pilus-structure: outside to inside
and top to bottom--a minireview. Gene 192:165-9.
Foxman, B. 2003. Epidemiology of urinary tract infections: incidence, morbidity,
and economic costs. Dis Mon 49:53-70.
Foxman, B. 1990. Recurring urinary tract infection: incidence and risk factors.
Am J Public Health 80:331-3.
Foxman, B., and P. Brown. 2003. Epidemiology of urinary tract infections:
transmission and risk factors, incidence, and costs. Infect Dis Clin North Am
17:227-41.
Francetic, O., D. Belin, C. Badaut, and A. P. Pugsley. 2000. Expression of the
endogenous type II secretion pathway in Escherichia coli leads to chitinase
secretion. Embo J 19:6697-703.
91
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.
42.
43.
Francetic, O., and A. P. Pugsley. 2005. Towards the identification of type II
secretion signals in a nonacylated variant of pullulanase from Klebsiella oxytoca.
J Bacteriol 187:7045-55.
Francis, K. P., J. Yu, C. Bellinger-Kawahara, D. Joh, M. J. Hawkinson, G.
Xiao, T. F. Purchio, M. G. Caparon, M. Lipsitch, and P. R. Contag. 2001.
Visualizing pneumococcal infections in the lungs of live mice using
bioluminescent Streptococcus pneumoniae transformed with a novel grampositive lux transposon. Infect Immun 69:3350-8.
Fullner, K. J., and J. J. Mekalanos. 1999. Genetic characterization of a new
type IV-A pilus gene cluster found in both classical and El Tor biotypes of Vibrio
cholerae. Infect Immun 67:1393-404.
Gerlach, R. G., S. U. Holzer, D. Jackel, and M. Hensel. 2007. Rapid
engineering of bacterial reporter gene fusions by using red recombination. Appl
Environ Microbiol 73:4234-42.
Giron, J. A., A. S. Ho, and G. K. Schoolnik. 1991. An inducible bundle-forming
pilus of enteropathogenic Escherichia coli. Science 254:710-3.
Giron, J. A., M. M. Levine, and J. B. Kaper. 1994. Longus: a long pilus
ultrastructure produced by human enterotoxigenic Escherichia coli. Mol
Microbiol 12:71-82.
Gregory, T. R. 2005. Synergy between sequence and size in large-scale
genomics. Nat Rev Genet 6:699-708.
Guignot, J., C. Chaplais, M. H. Coconnier-Polter, and A. L. Servin. 2007. The
secreted autotransporter toxin, Sat, functions as a virulence factor in Afa/Dr
diffusely adhering Escherichia coli by promoting lesions in tight junction of
polarized epithelial cells. Cell Microbiol 9:204-21.
Guyer, D. M., S. Radulovic, F. E. Jones, and H. L. Mobley. 2002. Sat, the
secreted autotransporter toxin of uropathogenic Escherichia coli, is a vacuolating
cytotoxin for bladder and kidney epithelial cells. Infect Immun 70:4539-46.
Hagberg, L., I. Engberg, R. Freter, J. Lam, S. Olling, and C. Svanborg Eden.
1983. Ascending, unobstructed urinary tract infection in mice caused by
pyelonephritogenic Escherichia coli of human origin. Infect Immun 40:273-83.
Hales, L. M., and H. A. Shuman. 1999. Legionella pneumophila contains a type
II general secretion pathway required for growth in amoebae as well as for
secretion of the Msp protease. Infect Immun 67:3662-6.
Haraoka, M., L. Hang, B. Frendeus, G. Godaly, M. Burdick, R. Strieter, and
C. Svanborg. 1999. Neutrophil recruitment and resistance to urinary tract
infection. J Infect Dis 180:1220-9.
Hedlund, M., M. Svensson, A. Nilsson, R. D. Duan, and C. Svanborg. 1996.
Role of the ceramide-signaling pathway in cytokine responses to P-fimbriated
Escherichia coli. J Exp Med 183:1037-44.
Herrington, D. A., R. H. Hall, G. Losonsky, J. J. Mekalanos, R. K. Taylor,
and M. M. Levine. 1988. Toxin, toxin-coregulated pili, and the toxR regulon are
essential for Vibrio cholerae pathogenesis in humans. J Exp Med 168:1487-92.
Hitziger, N., I. Dellacasa, B. Albiger, and A. Barragan. 2005. Dissemination of
Toxoplasma gondii to immunoprivileged organs and role of Toll/interleukin-1
92
44.
45.
46.
47.
48.
49.
50.
51.
52.
53.
54.
55.
56.
57.
58.
59.
receptor signalling for host resistance assessed by in vivo bioluminescence
imaging. Cell Microbiol 7:837-48.
Holechek, M. J. 2003. Glomerular filtration: an overview. Nephrol Nurs J
30:285-90; quiz 291-2.
Hunstad, D. A., S. S. Justice, C. S. Hung, S. R. Lauer, and S. J. Hultgren.
2005. Suppression of bladder epithelial cytokine responses by uropathogenic
Escherichia coli. Infect Immun 73:3999-4006.
Hutchens, M., and G. D. Luker. 2007. Applications of bioluminescence imaging
to the study of infectious diseases. Cell Microbiol.
Jiang, B., and S. P. Howard. 1992. The Aeromonas hydrophila exeE gene,
required both for protein secretion and normal outer membrane biogenesis, is a
member of a general secretion pathway. Mol Microbiol 6:1351-61.
Johnson, J. R. 1991. Virulence factors in Escherichia coli urinary tract infection.
Clin Microbiol Rev 4:80-128.
Johnson, J. R., and T. A. Russo. 2002. Extraintestinal pathogenic Escherichia
coli: "the other bad E coli". J Lab Clin Med 139:155-62.
Justice, S. S., C. Hung, J. A. Theriot, D. A. Fletcher, G. G. Anderson, M. J.
Footer, and S. J. Hultgren. 2004. Differentiation and developmental pathways
of uropathogenic Escherichia coli in urinary tract pathogenesis. Proc Natl Acad
Sci U S A 101:1333-8.
Justice, S. S., D. A. Hunstad, P. C. Seed, and S. J. Hultgren. 2006.
Filamentation by Escherichia coli subverts innate defenses during urinary tract
infection. Proc Natl Acad Sci U S A 103:19884-9.
Kaper, J. B., J. P. Nataro, and H. L. Mobley. 2004. Pathogenic Escherichia
coli. Nat Rev Microbiol 2:123-40.
Kau, A. L., D. A. Hunstad, and S. J. Hultgren. 2005. Interaction of
uropathogenic Escherichia coli with host uroepithelium. Curr Opin Microbiol
8:54-9.
Kornacker, M. G., and A. P. Pugsley. 1990. The normally periplasmic enzyme
beta-lactamase is specifically and efficiently translocated through the Escherichia
coli outer membrane when it is fused to the cell-surface enzyme pullulanase. Mol
Microbiol 4:1101-9.
Kostakioti, M., C. L. Newman, D. G. Thanassi, and C. Stathopoulos. 2005.
Mechanisms of protein export across the bacterial outer membrane. J Bacteriol
187:4306-14.
Landraud, L., C. Pulcini, P. Gounon, G. Flatau, P. Boquet, and E. Lemichez.
2004. E. coli CNF1 toxin: a two-in-one system for host-cell invasion. Int J Med
Microbiol 293:513-8.
Lane, M. C., and H. L. Mobley. 2007. Role of P-fimbrial-mediated adherence in
pyelonephritis and persistence of uropathogenic Escherichia coli (UPEC) in the
mammalian kidney. Kidney Int 72:19-25.
Luker, G. D., J. P. Bardill, J. L. Prior, C. M. Pica, D. Piwnica-Worms, and D.
A. Leib. 2002. Noninvasive bioluminescence imaging of herpes simplex virus
type 1 infection and therapy in living mice. J Virol 76:12149-61.
Maier, B., L. Potter, M. So, C. D. Long, H. S. Seifert, and M. P. Sheetz. 2002.
Single pilus motor forces exceed 100 pN. Proc Natl Acad Sci U S A 99:16012-7.
93
60.
61.
62.
63.
64.
65.
66.
67.
68.
69.
70.
71.
72.
73.
74.
Martin, P. R., M. Hobbs, P. D. Free, Y. Jeske, and J. S. Mattick. 1993.
Characterization of pilQ, a new gene required for the biogenesis of type 4
fimbriae in Pseudomonas aeruginosa. Mol Microbiol 9:857-68.
Martinez, J. J., M. A. Mulvey, J. D. Schilling, J. S. Pinkner, and S. J.
Hultgren. 2000. Type 1 pilus-mediated bacterial invasion of bladder epithelial
cells. Embo J 19:2803-12.
Mattick, J. S. 2002. Type IV pili and twitching motility. Annu Rev Microbiol
56:289-314.
Merz, A. J., and M. So. 2000. Interactions of pathogenic neisseriae with
epithelial cell membranes. Annu Rev Cell Dev Biol 16:423-57.
Merz, A. J., M. So, and M. P. Sheetz. 2000. Pilus retraction powers bacterial
twitching motility. Nature 407:98-102.
Mobley, H. L., D. M. Green, A. L. Trifillis, D. E. Johnson, G. R.
Chippendale, C. V. Lockatell, B. D. Jones, and J. W. Warren. 1990.
Pyelonephritogenic Escherichia coli and killing of cultured human renal proximal
tubular epithelial cells: role of hemolysin in some strains. Infect Immun 58:12819.
Mobley, H. L., K. G. Jarvis, J. P. Elwood, D. I. Whittle, C. V. Lockatell, R. G.
Russell, D. E. Johnson, M. S. Donnenberg, and J. W. Warren. 1993. Isogenic
P-fimbrial deletion mutants of pyelonephritogenic Escherichia coli: the role of
alpha Gal(1-4) beta Gal binding in virulence of a wild-type strain. Mol Microbiol
10:143-55.
Mulvey, M. A., Y. S. Lopez-Boado, C. L. Wilson, R. Roth, W. C. Parks, J.
Heuser, and S. J. Hultgren. 1998. Induction and evasion of host defenses by
type 1-piliated uropathogenic Escherichia coli. Science 282:1494-7.
Mulvey, M. A., J. D. Schilling, and S. J. Hultgren. 2001. Establishment of a
persistent Escherichia coli reservoir during the acute phase of a bladder infection.
Infect Immun 69:4572-9.
Mulvey, M. A., J. D. Schilling, J. J. Martinez, and S. J. Hultgren. 2000. Bad
bugs and beleaguered bladders: interplay between uropathogenic Escherichia coli
and innate host defenses. Proc Natl Acad Sci U S A 97:8829-35.
Mysorekar, I. U., and S. J. Hultgren. 2006. Mechanisms of uropathogenic
Escherichia coli persistence and eradication from the urinary tract. Proc Natl
Acad Sci U S A 103:14170-5.
Neish, A. S., A. T. Gewirtz, H. Zeng, A. N. Young, M. E. Hobert, V. Karmali,
A. S. Rao, and J. L. Madara. 2000. Prokaryotic regulation of epithelial
responses by inhibition of IkappaB-alpha ubiquitination. Science 289:1560-3.
Nunn, D. N., and S. Lory. 1991. Product of the Pseudomonas aeruginosa gene
pilD is a prepilin leader peptidase. Proc Natl Acad Sci U S A 88:3281-5.
Palmer, L. E., A. R. Pancetti, S. Greenberg, and J. B. Bliska. 1999. YopJ of
Yersinia spp. is sufficient to cause downregulation of multiple mitogen-activated
protein kinases in eukaryotic cells. Infect Immun 67:708-16.
Paranjpye, R. N., and M. S. Strom. 2005. A Vibrio vulnificus type IV pilin
contributes to biofilm formation, adherence to epithelial cells, and virulence.
Infect Immun 73:1411-22.
94
75.
76.
77.
78.
79.
80.
81.
82.
83.
84.
85.
86.
87.
88.
89.
90.
Parsons, C. L., C. W. Stauffer, and J. D. Schmidt. 1988. Reversible
inactivation of bladder surface glycosaminoglycan antibacterial activity by
protamine sulfate. Infect Immun 56:1341-3.
Peabody, C. R., Y. J. Chung, M. R. Yen, D. Vidal-Ingigliardi, A. P. Pugsley,
and M. H. Saier, Jr. 2003. Type II protein secretion and its relationship to
bacterial type IV pili and archaeal flagella. Microbiology 149:3051-72.
Possot, O. M., M. Gerard-Vincent, and A. P. Pugsley. 1999. Membrane
association and multimerization of secreton component pulC. J Bacteriol
181:4004-11.
Pugsley, A. P., O. Francetic, O. M. Possot, N. Sauvonnet, and K. R. Hardie.
1997. Recent progress and future directions in studies of the main terminal branch
of the general secretory pathway in Gram-negative bacteria--a review. Gene
192:13-9.
Py, B., L. Loiseau, and F. Barras. 2001. An inner membrane platform in the
type II secretion machinery of Gram-negative bacteria. EMBO Rep 2:244-8.
Roberts, J. A., B. I. Marklund, D. Ilver, D. Haslam, M. B. Kaack, G. Baskin,
M. Louis, R. Mollby, J. Winberg, and S. Normark. 1994. The Gal(alpha 14)Gal-specific tip adhesin of Escherichia coli P-fimbriae is needed for
pyelonephritis to occur in the normal urinary tract. Proc Natl Acad Sci U S A
91:11889-93.
Romih, R., P. Korosec, W. de Mello, Jr., and K. Jezernik. 2005.
Differentiation of epithelial cells in the urinary tract. Cell Tissue Res 320:259-68.
Ronald, A. 2002. The etiology of urinary tract infection: traditional and emerging
pathogens. Am J Med 113 Suppl 1A:14S-19S.
Russo, T. A., and J. R. Johnson. 2000. Proposal for a new inclusive designation
for extraintestinal pathogenic isolates of Escherichia coli: ExPEC. J Infect Dis
181:1753-4.
Sandkvist, M. 2001. Biology of type II secretion. Mol Microbiol 40:271-83.
Sandkvist, M. 2001. Type II secretion and pathogenesis. Infect Immun 69:352335.
Sandkvist, M., L. O. Michel, L. P. Hough, V. M. Morales, M. Bagdasarian,
M. Koomey, V. J. DiRita, and M. Bagdasarian. 1997. General secretion
pathway (eps) genes required for toxin secretion and outer membrane biogenesis
in Vibrio cholerae. J Bacteriol 179:6994-7003.
Sauvonnet, N., P. Gounon, and A. P. Pugsley. 2000. PpdD type IV pilin of
Escherichia coli K-12 can Be assembled into pili in Pseudomonas aeruginosa. J
Bacteriol 182:848-54.
Sauvonnet, N., and A. P. Pugsley. 1996. Identification of two regions of
Klebsiella oxytoca pullulanase that together are capable of promoting betalactamase secretion by the general secretory pathway. Mol Microbiol 22:1-7.
Sauvonnet, N., G. Vignon, A. P. Pugsley, and P. Gounon. 2000. Pilus
formation and protein secretion by the same machinery in Escherichia coli. Embo
J 19:2221-8.
Savage, D. C. 2001. Microbial biota of the human intestine: a tribute to some
pioneering scientists. Curr Issues Intest Microbiol 2:1-15.
95
91.
92.
93.
94.
95.
96.
97.
98.
99.
100.
101.
102.
103.
104.
Schilling, J. D., M. A. Mulvey, C. D. Vincent, R. G. Lorenz, and S. J.
Hultgren. 2001. Bacterial invasion augments epithelial cytokine responses to
Escherichia coli through a lipopolysaccharide-dependent mechanism. J Immunol
166:1148-55.
Strom, M. S., and S. Lory. 1993. Structure-function and biogenesis of the type
IV pili. Annu Rev Microbiol 47:565-96.
Strom, M. S., D. N. Nunn, and S. Lory. 1993. A single bifunctional enzyme,
PilD, catalyzes cleavage and N-methylation of proteins belonging to the type IV
pilin family. Proc Natl Acad Sci U S A 90:2404-8.
Tauschek, M., R. J. Gorrell, R. A. Strugnell, and R. M. Robins-Browne.
2002. Identification of a protein secretory pathway for the secretion of heat-labile
enterotoxin by an enterotoxigenic strain of Escherichia coli. Proc Natl Acad Sci U
S A 99:7066-71.
Thanabalu, T., E. Koronakis, C. Hughes, and V. Koronakis. 1998. Substrateinduced assembly of a contiguous channel for protein export from E.coli:
reversible bridging of an inner-membrane translocase to an outer membrane exit
pore. Embo J 17:6487-96.
Thanassi, D. G., and S. J. Hultgren. 2000. Multiple pathways allow protein
secretion across the bacterial outer membrane. Curr Opin Cell Biol 12:420-30.
Tonjum, T., and M. Koomey. 1997. The pilus colonization factor of pathogenic
neisserial species: organelle biogenesis and structure/function relationships--a
review. Gene 192:155-63.
Turner, L. R., J. C. Lara, D. N. Nunn, and S. Lory. 1993. Mutations in the
consensus ATP-binding sites of XcpR and PilB eliminate extracellular protein
secretion and pilus biogenesis in Pseudomonas aeruginosa. J Bacteriol 175:49629.
Uhlen, P., A. Laestadius, T. Jahnukainen, T. Soderblom, F. Backhed, G.
Celsi, H. Brismar, S. Normark, A. Aperia, and A. Richter-Dahlfors. 2000.
Alpha-haemolysin of uropathogenic E. coli induces Ca2+ oscillations in renal
epithelial cells. Nature 405:694-7.
Vignon, G., R. Kohler, E. Larquet, S. Giroux, M. C. Prevost, P. Roux, and A.
P. Pugsley. 2003. Type IV-like pili formed by the type II secreton: specificity,
composition, bundling, polar localization, and surface presentation of peptides. J
Bacteriol 185:3416-28.
Voulhoux, R., G. Ball, B. Ize, M. L. Vasil, A. Lazdunski, L. F. Wu, and A.
Filloux. 2001. Involvement of the twin-arginine translocation system in protein
secretion via the type II pathway. Embo J 20:6735-41.
Waldor, M. K., and J. J. Mekalanos. 1996. Lysogenic conversion by a
filamentous phage encoding cholera toxin. Science 272:1910-4.
Wall, D., and D. Kaiser. 1999. Type IV pili and cell motility. Mol Microbiol
32:1-10.
Welch, R. A., V. Burland, G. Plunkett, 3rd, P. Redford, P. Roesch, D. Rasko,
E. L. Buckles, S. R. Liou, A. Boutin, J. Hackett, D. Stroud, G. F. Mayhew, D.
J. Rose, S. Zhou, D. C. Schwartz, N. T. Perna, H. L. Mobley, M. S.
Donnenberg, and F. R. Blattner. 2002. Extensive mosaic structure revealed by
96
105.
106.
107.
108.
the complete genome sequence of uropathogenic Escherichia coli. Proc Natl Acad
Sci U S A 99:17020-4.
Whitchurch, C. B., M. Hobbs, S. P. Livingston, V. Krishnapillai, and J. S.
Mattick. 1991. Characterisation of a Pseudomonas aeruginosa twitching motility
gene and evidence for a specialised protein export system widespread in
eubacteria. Gene 101:33-44.
Whitchurch, C. B., and J. S. Mattick. 1994. Characterization of a gene, pilU,
required for twitching motility but not phage sensitivity in Pseudomonas
aeruginosa. Mol Microbiol 13:1079-91.
Zhang, D., G. Zhang, M. S. Hayden, M. B. Greenblatt, C. Bussey, R. A.
Flavell, and S. Ghosh. 2004. A toll-like receptor that prevents infection by
uropathogenic bacteria. Science 303:1522-6.
Zoutman, D. E., W. C. Hulbert, B. L. Pasloske, A. M. Joffe, K. Volpel, M. K.
Trebilcock, and W. Paranchych. 1991. The role of polar pili in the adherence of
Pseudomonas aeruginosa to injured canine tracheal cells: a semiquantitative
morphologic study. Scanning Microsc 5:109-24; discussion 124-6.
97